Field observations provide valuable data regarding nearshore tsunami impact, yet only in inundation areas where tsunami waves have already flooded. Therefore, tsunami modeling is essential to understand tsunami behavior and prepare for tsunami inundation. It is necessary that all numerical models used in tsunami emergency planning be subject to benchmark tests for validation and verification. This study focuses on two numerical codes, NAMI DANCE and FLOW-3D®, for validation and performance comparison. NAMI DANCE is an in-house tsunami numerical model developed by the Ocean Engineering Research Center of Middle East Technical University, Turkey and Laboratory of Special Research Bureau for Automation of Marine Research, Russia. FLOW-3D® is a general purpose computational fluid dynamics software, which was developed by scientists who pioneered in the design of the Volume-of-Fluid technique. The codes are validated and their performances are compared via analytical, experimental and field benchmark problems, which are documented in the ‘‘Proceedings and Results of the 2011 National Tsunami Hazard Mitigation Program (NTHMP) Model Benchmarking Workshop’’ and the ‘‘Proceedings and Results of the NTHMP 2015 Tsunami Current Modeling Workshop”. The variations between the numerical solutions of these two models are evaluated through statistical error analysis.
현장 관찰은 연안 쓰나미 영향에 관한 귀중한 데이터를 제공하지만 쓰나미 파도가 이미 범람한 침수 지역에서만 가능합니다. 따라서 쓰나미 모델링은 쓰나미 행동을 이해하고 쓰나미 범람에 대비하는 데 필수적입니다.
쓰나미 비상 계획에 사용되는 모든 수치 모델은 검증 및 검증을 위한 벤치마크 테스트를 받아야 합니다. 이 연구는 검증 및 성능 비교를 위해 NAMI DANCE 및 FLOW-3D®의 두 가지 숫자 코드에 중점을 둡니다.
NAMI DANCE는 터키 중동 기술 대학의 해양 공학 연구 센터와 러시아 해양 연구 자동화를 위한 특별 조사국 연구소에서 개발한 사내 쓰나미 수치 모델입니다. FLOW-3D®는 Volume-of-Fluid 기술의 설계를 개척한 과학자들이 개발한 범용 전산 유체 역학 소프트웨어입니다.
코드의 유효성이 검증되고 분석, 실험 및 현장 벤치마크 문제를 통해 코드의 성능이 비교되며, 이는 ‘2011년 NTHMP(National Tsunami Hazard Mitigation Program) 모델 벤치마킹 워크숍의 절차 및 결과’와 ”절차 및 NTHMP 2015 쓰나미 현재 모델링 워크숍 결과”. 이 두 모델의 수치 해 사이의 변동은 통계적 오류 분석을 통해 평가됩니다.
Apotsos, A., Buckley, M., Gelfenbaum, G., Jafe, B., & Vatvani, D. (2011). Nearshore tsunami inundation and sediment transport modeling: towards model validation and application. Pure and Applied Geophysics,168(11), 2097–2119. https://doi.org/10.1007/s00024-011-0291-5.ArticleGoogle Scholar
Barberopoulou, A., Legg, M. R., & Gica, E. (2015). Time evolution of man-made harbor modifications in San Diego: effects on Tsunamis. Journal of Marine Science and Engineering,3, 1382–1403.ArticleGoogle Scholar
Basu, D., Green, S., Das, K., Janetzke, R. and Stamatakos, J. (2009). Numerical Simulation of Surface Waves Generated by a Subaerial Landslide at Lituya Bay, Alaska. Proceedings of 28th International Conference on Ocean, Offshore and Arctic Engineering. Honolulu, Hawaii, USA.
Briggs, M. J., Synolakis, C. E., Harkins, G. S., & Green, D. R. (1995). Laboratory experiments of tsunami run-up on a circular island. Pure and Applied Geophysics,144(3/4), 569–593.ArticleGoogle Scholar
Cheung, K. F., Bai, Y., & Yamazaki, Y. (2013). Surges around the Hawaiian Islands from the 2011 Tohoku Tsunami. Journal of Geophysical Research: Oceans,118, 5703–5719. https://doi.org/10.1002/jgrc.20413.Google Scholar
Choi, B. H., Dong, C. K., Pelinovsky, E., & Woo, S. B. (2007). Three-dimensional Simulation of Tsunami Run-up Around Conical Island. Coastal Engineering,54, 618–629.ArticleGoogle Scholar
Cox, D., T. Tomita, P. Lynett, R.A., Holman. (2008). Tsunami Inundation with Macroroughness in the Constructed Environment. Proceedings of 31st International Conference on Coastal Engineering, ASCE, pp. 1421–1432.
Flow Science. (2002). FLOW-3D User’s Manual.
Hirt, C. W., & Nichols, B. D. (1981). Volume of fluid (VOF) method for the dynamics of free boundaries. Journal of Computational Physics,39, 201–225.ArticleGoogle Scholar
Horrillo, J., Grilli, S. T., Nicolsky, D., Roeber, V., & Zang, J. (2015). Performance benchmarking Tsunami models for NTHMP’s inundation mapping activities. Pure and Applied Geophysics,172, 869–884.ArticleGoogle Scholar
Kim, K. O., Kim, D. C., Choi, B.-H., Jung, T. K., Yuk, J. H., & Pelinovsky, E. (2015). The role of diffraction effects in extreme run-up inundation at Okushiri Island due to 1993 Tsunami. Natural Hazards and Earth Systems Sciences,15, 747–755.ArticleGoogle Scholar
Liu, P. L.-F. (1994). Model equations for wave propagations from deep to shallow water. (P.-F. Liu, Ed.) Advances in Coastal and Ocean Engineering, 1, 125–158.
Liu, P. L.-F., Yeh, H., & Synolakis, C. E. (2008). Advanced numerical models for simulating Tsunami waves and run-up. Advances in Coastal and Ocean Engineering,10, 344.Google Scholar
Lynett, P. J., Gately, K., Wilson, R., Montoya, L., Arcas, D., Aytore, B., et al. (2017). Inter-model analysis of Tsunami-induced coastal currents. Ocean Modelling,114, 14–32.ArticleGoogle Scholar
Lynett, P. J., Wu, T.-R., & Liu, P. L.-F. (2002). Modeling wave run-up with depth-integrated equations. Coastal Engineering,46(2), 89–107.ArticleGoogle Scholar
Macias, J., Castro, M. J., Ortega, S., Escalante, C., & Gonzalez-Vida, J. M. (2017). Performance benchmarking of Tsunami-HySEA model for nthmp’s inundation mapping activities. Pure and Applied Geophysics,174, 3147–3183.ArticleGoogle Scholar
Matsuyama, M., & Tanaka, H. (2001). An experimental study of the highest run-up height in the 1993 Hokkaidō Nansei-Oki Earthquake Tsunami. Proceedings of ITS,2001, 879–889.Google Scholar
National Tsunami Hazard Mitigation Program. 2012. Proceedings and Results of the 2011 NTHMP Model Benchmarking Workshop. Boulder: U.S. Department of Commerce/NOAA/NTHMP; (NOAA Special Report). p. 436.
National Tsunami Hazard Mitigation Program. (2017). Proceedings and Results of the National Tsunami Hazard Mitigation Program 2015 Tsunami Current Modeling Workshop, February 9-10, 2015, Portland, Oregon: compiled by Patrick Lynett and Rick Wilson, p 194.
Necmioglu, O., & Ozel, N. M. (2014). An earthquake source sensitivity analysis for Tsunami propagation in the Eastern Mediterranean. Oceanography,27(2), 76–85.ArticleGoogle Scholar
Nichols, B.D. and Hirt, C.W. (1975). Methods for Calculating Multi-Dimensional, Transient Free Surface Flows Past Bodies. Proceedings of 1st International Conference Num. Ship Hydrodynamics. Gaithersburg.
Nicolsky, D. J., Suleimani, E. N., & Hansen, R. A. (2011). Validation and verification of a numerical model for Tsunami propagation and run-up. Pure and Applied Geophysics,168(6), 1199–1222.ArticleGoogle Scholar
Park, H., Cox, D. T., Lynett, P. J., Wiebe, D. M., & Shin, S. (2013). Tsunami inundation modeling in constructed environments: a physical and numerical comparison of free-surface elevation, velocity, and momentum flux. Coastal Engineering,79, 9–21.ArticleGoogle Scholar
Patel, V. M., Dholakia, M. B., & Singh, A. P. (2016). Emergency preparedness in the case of Makran Tsunami: a case study on Tsunami risk visualization for the Western Parts of Gujarat, India. Geomatics Natural Hazard and Risk,7(2), 826–842.ArticleGoogle Scholar
Pelinovsky, E., Kim, D.-C., Kim, K.-O., & Choi, B.-H. (2013). Three-dimensional simulation of extreme run-up heights during the 2004 Indonesian and 2011 Japanese Tsunamis. Vienna: EGU General Assembly.Google Scholar
Rueben, M., Holman, R., Cox, D., Shin, S., Killian, J., & Stanley, J. (2011). Optical measurements of Tsunami inundation through an urban waterfront modeled in a large-scale laboratory basin. Coastal Engineering,58, 229–238.ArticleGoogle Scholar
Shuto, N. (1991). Numerical simulation of Tsunamis—its present and near future. Natural Hazards,4, 171–191.ArticleGoogle Scholar
Synolakis, C. E. (1986). The run-up of long waves. Ph.D. Thesis. California Institute of Technology, Pasadena, California.
Synolakis, C. E., Bernard, E. N., Titov, V. V., Kanoglu, U. & Gonzalez, F. (2007). Standards, criteria, and procedures for NOAA evaluation of Tsunami Numerical Models. 55 p. Seattle, Washington: NOAA OAR Special Report, Contribution No 3053, NOAA/OAR/PMEL.
Synolakis, C. E., Bernard, E. N., Titov, V. V., Kanoglu, U., & Gonzalez, F. I. (2008). Validation and verification of Tsunami numerical models. Pure and Applied Geophysics,165, 2197–2228.ArticleGoogle Scholar
Tolkova, E. (2014). Land-water boundary treatment for a tsunami model with dimensional splitting. Pure and Applied Geophysics,171(9), 2289–2314.ArticleGoogle Scholar
Velioglu, D. (2017). Advanced two- and three-dimensional Tsunami models: benchmarking and validation. Ph.D. Thesis. Middle East Technical University, Ankara.
Velioglu, D., Kian, R., Yalciner, A.C. and Zaytsev, A. (2016). Performance assessment of NAMI DANCE in Tsunami evolution and currents using a benchmark problem. (R. Signell, Ed.) J. Mar. Sci. Eng., 4(3), 49.
Wu, T. (2001). A unified theory for modeling water waves. Advances in Applied Mechanics,37, 1–88.ArticleGoogle Scholar
Wu, N.-J., Hsiao, S.-C., Chen, H.-H., & Yang, R.-Y. (2016). The study on solitary waves generated by a piston-type wave maker. Ocean Engineering,117, 114–129.ArticleGoogle Scholar
Yalciner, A. C., Dogan, P. and Sukru. E. (2005). December 26 2004, Indian Ocean Tsunami Field Survey, North of Sumatra Island. UNESCO.
Yalciner, A. C., Gülkan, P., Dilmen, I., Aytore, B., Ayca, A., Insel, I., et al. (2014). Evaluation of Tsunami scenarios For Western Peloponnese, Greece. Bollettino di Geofisica Teorica ed Applicata,55, 485–500.Google Scholar
Yen, B. C. (1991). Hydraulic resistance in open channels. In B. C. Yen (Ed.), Channel flow resistance: centennial of manning’s formula (pp. 1–135). Highlands Ranch: Water Resource Publications.Google Scholar
Zaitsev, A. I., Kovalev, D. P., Kurkin, A. A., Levin, B. V., Pelinovskii, E. N., Chernov, A. G., et al. (2009). The Tsunami on Sakhalin on August 2, 2007: mareograph evidence and numerical simulation. Tikhookeanskaya Geologiya,28, 30–35.Google Scholar
The authors wish to thank Dr. Andrey Zaytsev due to his undeniable contributions to the development of in-house numerical model, NAMI DANCE. The Turkish branch of Flow Science, Inc. is also acknowledged. Finally, the National Tsunami Hazard Mitigation Program (NTHMP), who provided most of the benchmark data, is appreciated. This research did not receive any specific grant from funding agencies in the public, commercial, or not-for-profit sectors.
Author information
Author notes
Deniz Velioglu SogutPresent address: 1212 Computer Science, Department of Civil Engineering, Stony Brook University, Stony Brook, NY, 11794, USA
Authors and Affiliations
Middle East Technical University, 06800, Ankara, TurkeyDeniz Velioglu Sogut & Ahmet Cevdet Yalciner
Velioglu Sogut, D., Yalciner, A.C. Performance Comparison of NAMI DANCE and FLOW-3D® Models in Tsunami Propagation, Inundation and Currents using NTHMP Benchmark Problems. Pure Appl. Geophys.176, 3115–3153 (2019). https://doi.org/10.1007/s00024-018-1907-9
The hydrodynamics of coral reefs strongly influences their biological functioning, impacting processes such as nutrient availability and uptake, recruitment success and bleaching. For example, coral reefs located in oligotrophic regions depend on upwelling for nutrient supply. Coral reefs at Sodwana Bay, located on the east coast of South Africa, are an example of high latitude marginal reefs. These reefs are subjected to complex hydrodynamic forcings due to the interaction between the strong Agulhas current and the highly variable topography of the region. In this study, we explore the reef scale hydrodynamics resulting from the bathymetry for two steady current scenarios at Two-Mile Reef (TMR) using a combination of field data and numerical simulations. The influence of tides or waves was not considered for this study as well as reef-scale roughness. Tilt current meters with onboard temperature sensors were deployed at selected locations within TMR. We used field observations to identify the dominant flow conditions on the reef for numerical simulations that focused on the hydrodynamics driven by mean currents. During the field campaign, southerly currents were the predominant flow feature with occasional flow reversals to the north. Northerly currents were associated with greater variability towards the southern end of TMR. Numerical simulations showed that Jesser Point was central to the development of flow features for both the northerly and southerly current scenarios. High current variability in the south of TMR during reverse currents is related to the formation of Kelvin-Helmholtz type shear instabilities along the outer edge of an eddy formed north of Jesser Point. Furthermore, downward vertical velocities were computed along the offshore shelf at TMR during southerly currents. Current reversals caused a change in vertical velocities to an upward direction due to the orientation of the bathymetry relative to flow directions.
Highlights
A predominant southerly current was measured at Two-Mile Reef with occasional reversals towards the north.
Field observations indicated that northerly currents are spatially varied along Two-Mile Reef.
Simulation of reverse currents show the formation of a separated flow due to interaction with Jesser Point with Kelvin–Helmholtz type shear instabilities along the seaward edge.
지금까지 Sodwana Bay에서 자세한 암초 규모 유체 역학을 모델링하려는 시도는 없었습니다. 이러한 모델의 결과는 규모가 있는 산호초 사이의 흐름이 산호초 건강에 어떤 영향을 미치는지 탐색하는 데 사용할 수 있습니다. 이 연구에서는 Sodwana Bay의 유체역학을 탐색하는 데 사용할 수 있는 LES 모델을 개발하기 위한 단계별 접근 방식을 구현합니다. 여기서 우리는 이 초기 단계에서 파도와 조수의 영향을 배제하면서 Agulhas 해류의 유체역학에 초점을 맞춥니다. 이 접근법은 흐름의 첫 번째 LES를 제시하고 Sodwana Bay의 산호초에서 혼합함으로써 향후 연구의 기초를 제공합니다.
Booij N, Ris RC, Holthuijsen LH (1999) A third-generation wave model for coastal regions: 1. Model description and validation. J Geophys Res Ocean 104(C4):7649–7666. https://doi.org/10.1029/98JC02622ArticleGoogle Scholar
Celliers L, Schleyer MH (2002) Coral bleaching on high-latitude marginal reefs at Sodwana Bay, South Africa. Mar Pollut Bull 44:1380–1387ArticleGoogle Scholar
Chen SC (2018) Performance assessment of FLOW-3D and XFlow in the numerical modelling of fish-bone type fishway hydraulics https://doi.org/10.15142/T3HH1J
Flow Science Inc (2018) FLOW-3D, Version 12.0 Users Manual. Santa Fe, NM, https://www.flow3d.com/
Flow Science Inc (2019) FLOW-3D, Version 12.0 [Computer Software]. Santa Fe, NM, https://www.flow3d.com/
Franco A, Moernaut J, Schneider-Muntau B, Strasser M, Gems B (2020) The 1958 Lituya Bay tsunami – pre-event bathymetry reconstruction and 3D numerical modelling utilising the computational fluid dynamics software Flow-3D. Nat Hazards Earth Syst Sci 20(8):2255–2279ArticleGoogle Scholar
Fringer OB, Gerritsen M, Street RL (2006) An unstructured-grid, finite-volume, nonhydrostatic, parallel coastal ocean simulator. Ocean Model 14(3):139–173ArticleGoogle Scholar
Hirt CW, Sicilian JM (1985) A porosity technique for the definition of obstacles in rectangular cell meshes. In: Proceedings of 4th International Conference on Ship Hydrodynamics https://ci.nii.ac.jp/naid/10009570543/en/
Hocker LO, Hruska MA (2004) Interleaving synchronous data and asynchronous data in a single data storage file
Lim A, Wheeler AJ, Price DM, O’Reilly L, Harris K, Conti L (2020) Influence of benthic currents on cold-water coral habitats: a combined benthic monitoring and 3D photogrammetric investigation. Sci Rep 10(1):19433. https://doi.org/10.1038/s41598-020-76446-yArticleGoogle Scholar
Morris T (2009) Physical oceanography of Sodwana Bay and its effect on larval transport and coral bleaching. PhD thesis, Cape Peninsula University of Technology
Pope SB (2001) Turbulent flows. Cambridge University Press, CambridgeGoogle Scholar
Porter SN (2009) Biogeography and potential factors regulating shallow subtidal reef communities in the Western Indian Ocean. PhD thesis, University of Cape Town
Porter SN, Schleyer MH (2019) Environmental variation and how its spatial structure influences the cross-shelf distribution of high-latitude coral communities in South Africa. Diversity. https://doi.org/10.3390/d11040057ArticleGoogle Scholar
Ramsay PJ, Mason TR (1990) Development of a type zoning model for Zululand coral reefs, Sodwana Bay, South Africa. J Coastal Res 6(4):829–852Google Scholar
Roberts H, Richardson J, Lagumbay R, Meselhe E, Ma Y (2013) Hydrodynamic and sediment transport modeling using FLOW-3D for siting and optimization of the LCA medium diversion at white ditch hydrodynamic and sediment transport modeling using FLOW-3D for siting and optimization of the LCA medium diversion at white D (December)
Roberts MJ, Ribbink AJ, Morris T, Berg MAVD, Engelbrecht DC, Harding RT (2006) Oceanographic environment of the Sodwana Bay coelacanths (Latimeria chalumnae), South Africa: coelacanth research. South Afr J Sci 102(9):435–443Google Scholar
Rogers JS, Monismith SG, Koweek DA, Torres WI, Dunbar RB (2016) Thermodynamics and hydrodynamics in an atoll reef system and their influence on coral cover. Limnol Oceanogr 61(6):2191–2206. https://doi.org/10.1002/lno.10365ArticleGoogle Scholar
Schleyer MH, Celliers L (2003) Coral dominance at the reef-sediment interface in marginal coral communities at Sodwana Bay, South Africa. Mar Freshw Res 54(8):967–972. https://doi.org/10.1071/MF02049ArticleGoogle Scholar
Schleyer MH, Porter SN (2018) Chapter One – drivers of soft and stony coral community distribution on the high-latitude coral reefs of South Africa. advances in marine biology, vol 80, Academic Press, pp 1–55, https://doi.org/10.1016/bs.amb.2018.09.001
Sebens KP, Grace SP, Helmuth B, Maney EJ Jr, Miles JS (1998) Water flow and prey capture by three scleractinian corals, Madracis mirabilis, Montastrea cavernosa and Porites porites, in a field enclosure. Mar Biol 131(2):347–360ArticleGoogle Scholar
Smagorinsky J (1963) General circulation experiments with the primitive equations. Mon Weather Rev 91(3):99–164ArticleGoogle Scholar
Stocking J, Laforsch C, Sigl R, Reidenbach M (2018) The role of turbulent hydrodynamics and surface morphology on heat and mass transfer in corals. J R Soc Interface 15:20180448. https://doi.org/10.1098/rsif.2018.0448ArticleGoogle Scholar
Wyatt ASJ, Lowe RJ, Humphries S, Waite AM (2010) Particulate nutrient fluxes over a fringing coral reef: relevant scales of phytoplankton production and mechanisms of supply. Mar Ecol Prog Ser 405:113–130ArticleGoogle Scholar
NesreenTahabMaged M.El-FekyaAtef A.El-SaiadaIsmailFathya aDepartment of Water and Water Structures Engineering, Faculty of Engineering, Zagazig University, Zagazig 44519, Egypt bLab Manager, Faculty of Engineering, Zagazig University, Zagazig 44519, Egypt
Abstract
횡단 구조물을 통한 막힘은 안정성을 위협하는 위험한 문제 중 하나입니다. 암거의 막힘 형상 및 하류 세굴 특성에 미치는 영향에 관한 연구는 거의 없습니다.
이 연구의 목적은 수면과 세굴 모두에서 상자 암거를 통한 막힘의 작용을 수치적으로 논의하는 것입니다. 이를 위해 FLOW 3D v11.1.0을 사용하여 퇴적물 수송 모델을 조사했습니다.
상자 암거를 통한 다양한 차단 비율이 연구되었습니다. FLOW 3D 모델은 실험 데이터로 보정되었습니다. 결과는 FLOW 3D 프로그램이 세굴 다운스트림 상자 암거를 정확하게 시뮬레이션할 수 있음을 나타냅니다.
막힌 경우에 대한 속도 분포, 최대 세굴 깊이 및 수심을 플롯하고 비차단된 사례(기본 사례)와 비교했습니다.
그 결과 암거 높이의 70% 차단율은 상류의 수심을 암거 높이의 2.3배 증가시키고 평균 유속은 기본 경우보다 3배 더 증가시키는 것으로 입증되었다. 막힘 비율의 함수로 상대 최대 세굴 깊이를 추정하는 방정식이 만들어졌습니다.
Blockage through crossing structures is one of the dangerous problems that threaten its stability. There are few researches concerned with blockage shape in culverts and its effect on characteristics of scour downstream it.
The study’s purpose is to discuss the action of blockage through box culvert on both water surface and scour numerically. A sediment transport model has been investigated for this purpose using FLOW 3D v11.1.0. Different ratios of blockage through box culvert have been studied. The FLOW 3D model was calibrated with experimental data.
The results present that the FLOW 3D program was capable to simulate accurately the scour downstream box culvert. The velocity distribution, maximum scour depth and water depths for blocked cases have been plotted and compared with the non-blocked case (base case).
The results proved that the blockage ratio 70% of culvert height makes the water depth upstream increases by 2.3 times of culvert height and mean velocity increases by 3 times more than in the base case. An equation has been created to estimate the relative maximum scour depth as a function of blockage ratio.
1. Introduction
Local scour is the removal of granular bed material by the action of hydrodynamic forces. As the depth of scour hole increases, the stability of the foundation of the structure may be endangered, with a consequent risk of damage and failure [1]. So the prediction and control of scour is considered to be very important for protecting the water structures from failure. Most previous studies were designed to study the different factors that impact on scour and their relationship with scour hole dimensions like fluid characteristics, flow conditions, bed properties, and culvert geometry. Many previous researches studied the effect of flow rate on scour hole by information Froude number or modified Froude number [2], [3], [4], [5], [6]. Cesar Mendoza [6] found a good correlation between the scour depth and the discharge Intensity (Qg−.5D−2.5). Breusers and Raudkiv [7] used shear velocity in the outlet-scour prediction procedure. Ali and Lim [8] used the densimetric Froude number in estimation of the scour depth [1], [8], [9], [10], [11], [12], [13], [14]. “The densimetric Froude number presents the ratio of the tractive force on sediment particle to the submerged specific weight of the sediment” [15](1)Fd=uρsρ-1gD50
Ali and Lim [8] pointed to the consequence of tailwater depth on scour behavior [1], [2], [8], [13]. Abida and Townsend [2] indicated that the maximum depth of local scour downstream culvert was varying with the tailwater depth in three ways: first, for very shallow tailwater depths, local scouring decreases with a decrease in tailwater depth; second, when the ratio of tailwater depth to culvert height ranged between 0.2 and 0.7, the scour depth increases with decreasing tailwater depth; and third for a submerged outlet condition. The tailwater depth has only a marginal effect on the maximum depth of scour [2]. Ruff et al. [16] observed that for materials having similar mean grain sizes (d50) but different standard deviations (σ). As (σ) increased, the maximum scour hole depth decreased. Abt et al. [4] mentioned to role of soil type of maximum scour depth. It was noticed that local scour was more dangerous for uniform sands than for well-graded mixtures [1], [2], [4], [9], [17], [18]. Abt et al [3], [19] studied the culvert shape effect on scour hole. The results evidenced that the culvert shape has a limited effect on outlet scour. Under equivalent discharge conditions, it was noted that a square culvert with height equal to the diameter of a circular culvert would reduce scour [16], [20]. The scour hole dimension was also effected by the culvert slope. Abt et al. [3], [21] showed that the culvert slope is a key element in estimating the culvert flow velocity, the discharge capacity, and sediment transport capability. Abt et al. [21], [22] tested experimentally culvert drop height effect on maximum scour depth. It was observed that as the drop height was increasing, the depth of scour was also increasing. From the previous studies, it could have noticed that the most scour prediction formula downstream unblocked culvert was the function of densimetric Froude number, soil properties (d50, σ), tailwater depth and culvert opening size. Blockage is the phenomenon of plugging water structures due to the movement of water flow loaded with sediment and debris. Water structures blockage has a bad effect on water flow where it causes increasing of upstream water level that may cause flooding around the structure and increase of scour rate downstream structures [23], [24]. The blockage phenomenon through was studied experimentally and numerical [15], [25], [26], [27], [28], [29], [30], [31], [32], [33]. Jaeger and Lucke [33] studied the debris transport behavior in a natural channel in Australia. Froude number scale model of an existing culvert was used. It was noticed that through rainfall event, the mobility of debris was impressed by stream shape (depth and width). The condition of the vegetation (size and quantities) through the catchment area was the main factor in debris transport. Rigby et al. [26] reported that steep slope was increasing the ability to mobilize debris that form field data of blocked culverts and bridges during a storm in Wollongong city.
Streftaris et al. [32] studied the probability of screen blockage by debris at trash screens through a numerical model to relate between the blockage probability and nature of the area around. Recently, many commercial computational fluid programs (CFD) such as SSIIM, Fluent, and FLOW 3D are used in the analysis of the scour process. Scour and sediment transport numerical model need to validate by using experimental data or field data [34], [35], [36], [37], [38]. Epely-Chauvin et al. [36] investigated numerically the effect of a series of parallel spur diked. The experimental data were compared by SSIIM and FLOW 3D program. It was found that the accuracy of calibrated FLOW 3D model was better than SSIIM model. Nielsen et al. [35] used the physical model and FLOW 3D model to analyze the scour process around the pile. The soil around the pile was uniform coarse stones in the physical models that were simulated by regular spheres, porous media, and a mixture of them. The calibrated porous media model can be used to determine the bed shear stress. In partially blocked culverts, there aren’t many studies that explain the blockage impact on scour dimensions. Sorourian et al. [14], [15] studied the effect of inlet partial blockage on scour characteristics downstream box culvert. It resulted that the partial blockage at the culvert inlet could be the main factor in estimating the depth of scour. So, this study is aiming to investigate the effects of blockage through a box culvert on flow and scour characteristics by different blockage ratios and compares the results with a non-blocked case. Create a dimensionless equation relates the blockage ratio of the culvert with scour characteristics downstream culvert.
2. Experimental data
The experimental work of the study was conducted in the Hydraulics and Water Engineering Laboratory, Faculty of Engineering, Zagazig University, Egypt. The flume had a rectangular cross-section of 66 cm width, 65.5 cm depth, and 16.2 m long. A rectangular culvert was built with 0.2 m width, 0.2 m height and 3.00 m long with θ = 25° gradually outlet and 0.8 m fixed apron. The model was located on the mid-point of the channel. The sediment part was extended for a distance 2.20 m with 0.66 m width and 0.20 m depth of coarse sand with specific weight 1.60 kg/cm3, d50 = 2.75 mm and σ (d90/d50) = 1.50. The particle size distribution was as shown in Fig. 1. The experimental model was tested for different inlet flow (Q) of 25, 30, 34, 40 l/s for different submerged ratio (S) of 1.25, 1.50, 1.75.
3. Dimensional analysis
A dimensional analysis has been used to reduce the number of variables which affecting on the scour pattern downstream partial blocked culvert. The main factors affecting the maximum scour depth are:(2)ds=f(b.h.L.hb.lb.Q.ud.hu.hd.D50.ρ.ρs.g.ls.dd.ld)
Fig. 2 shows a definition sketch of the experimental model. The maximum scour depth can be written in a dimensionless form as:(3)dsh=f(B.Fd.S)where the ds/h is the relative maximum scour depth.
4. Numerical work
The FLOW 3D is (CFD) program used by many researchers and appeared high accuracy in solving hydrodynamic and sediment transport models in the three dimensions. Numerical simulation with FLOW 3D was performed to study the impacts of blockage ratio through box culvert on shear stress, velocity distribution and the sediment transport in terms of the hydrodynamic features (water surface, velocity and shear stress) and morphological parameters (scour depth and sizes) conditions in accurately and efficiently. The renormalization group (RNG) turbulence model was selected due to its high ability to predict the velocity profiles and turbulent kinetic energy for the flow through culvert [39]. The one-fluid incompressible mode was used to simulate the water surface. Volume of fluid (VOF) method was employed in FLOW 3D to tracks a liquid interface through arbitrary deformations and apply the correct boundary conditions at the interface [40].1.
Governing equations
Three-dimensional Reynolds-averaged Navier Stokes (RANS) equation was applied for incompressible viscous fluid motion. The continuity equation is as following:(4)VF∂ρ∂t+∂∂xρuAx+∂∂yρvAy+∂∂zρwAz=RDIF(5)∂u∂t+1VFuAx∂u∂x+vAy∂u∂y+ωAz∂u∂z=-1ρ∂P∂x+Gx+fx(6)∂v∂t+1VFuAx∂v∂x+vAy∂v∂y+ωAz∂v∂z=-1ρ∂P∂y+Gy+fy(7)∂ω∂t+1VFuAx∂ω∂x+vAy∂ω∂y+ωAz∂ω∂z=-1ρ∂P∂z+Gz+fz
ρ is the fluid density,
VF is the volume fraction,
(x,y,z) is the Cartesian coordinates,
(u,v,w) are the velocity components,
(Ax,Ay,Az) are the area fractions and
RDIF is the turbulent diffusion.
P is the average hydrodynamic pressure,
(Gx, Gy, Gz) are the body accelerations and
(fx, fy, fz) are the viscous accelerations.
The motion of sediment transport (suspended, settling, entrainment, bed load) is estimated by predicting the erosion, advection and deposition process as presented in [41].
The critical shields parameter is (θcr) is defined as the critical shear stress τcr at which sediments begin to move on a flat and horizontal bed [41]:(8)θcr=τcrgd50(ρs-ρ)
The Soulsby–Whitehouse [42] is used to predict the critical shields parameter as:(9)θcr=0.31+1.2d∗+0.0551-e(-0.02d∗)(10)d∗=d50g(Gs-1ν3where:
d* is the dimensionless grain size
Gs is specific weight (Gs = ρs/ρ)
The entrainment coefficient (0.005) was used to scale the scour rates and fit the experimental data. The settling velocity controls the Soulsby deposition equation. The volumetric sediment transport rate per width of the bed is calculated using Van Rijn [43].2.
Meshing and geometry of model
After many trials, it was found that the uniform cell size with 0.03 m cell size is the closest to the experimental results and takes less time. As shown in Fig. 3. In x-direction, the total model length in this direction is 700 cm with mesh planes at −100, 0, 300, 380 and 600 cm respectively from the origin point, in y-direction, the total model length in this direction is 66 cm at distances 0, 23, 43 and 66 cm respectively from the origin point. In z-direction, the total model length in this direction is 120 cm. with mesh planes at −20, 0, 20 and 100 cm respectively.3.
Boundary condition
As shown in Fig. 4, the boundary conditions of the model have been defined to simulate the experimental flow conditions accurately. The upstream boundary was defined as the volume flow rate with a different flow rate. The downstream boundary was defined as specific pressure with different fluid elevation. Both of the right side, the left side, and the bottom boundary were defined as a wall. The top boundary defined as specified pressure with pressure value equals zero.
5. Validation of experimental results and numerical results
The experimental results investigated the flow and scour characteristics downstream culvert due to different flow conditions. The measured value of maximum scour depth is compared with the simulated depth from FLOW 3D model as shown in Fig. 5. The scour results show that the simulated results from the numerical model is quite close to the experimental results with an average error of 3.6%. The water depths in numerical model results is so close to the experimental results as shown in Fig. 6 where the experiment and numerical results are compared at different submerged ratios and flow rates. The results appear maximum error percentage in water depths upstream and downstream the culvert is about 2.37%. This indicated that the FLOW 3D is efficient for the prediction of maximum scour depth and the flow depths downstream box culvert.
6. Computation time
The run time was chosen according to reaching to the stability limit. Hydraulic stability was achieved after 50 s, where the scour development may still go on. For run 1, the numerical simulation was run for 1000 s as shown in Fig. 7 where it mostly reached to scour stability at 800 s. The simulation time was taken 500 s at about 95% of scour stability.
7. Analysis and discussions
Fig. 8 shows the study sections where sec 1 represents to upstream section, sec2 represents to inside section and sec3 represents to downstream stream section. Table 1 indicates the scour hole dimensions at different blockage case. The symbol (B) represents to blockage and the number points to blockage ratio. B0 case signifies to the non-blocked case, B30 is that blockage height is 30% to the culvert height and so on.
Table 1. The scour results of different blockage ratio.
Case
hb cm
B = hb/h
Q lit/s
S
Fd
d50 mm
ds/h measured
ls/h
dd/h
ld/h
ds/h estimated
B0
0
0
35
1.26
1.69
2.5
0.58
1.50
0.27
5.00
0.46
B30
6
0.30
35
1.26
1.68
2.5
0.48
1.25
0.27
4.25
0.40
B50
10
0.50
35
1.22
1.74
2.5
0.45
1.10
0.24
4.00
0.37
B70
14
0.70
35
1.23
1.73
2.5
0.43
1.50
0.16
5.50
0.33
7.1. Scour hole geometry
The scour hole geometry mainly depends on the properties of soil of the bed downstream the fixed apron. From Table 1, the results show that the maximum scour depth in B0 case is about 0.58 of culvert height while the maximum deposition in B0 is 0.27 culvert height. There is a symmetric scour hole as shown in Fig. 9 in B0 case. An asymmetric scour hole is created in B50 and B70 due to turbulences that causes the deviation of the jet direction from the center of the flume where appear in Fig. 11 and Fig. 19.
7.2. Flow water surface
Fig. 10 presents the relative free surface water (hw/h) along the x-direction at center of the box culvert. From the mention Figure, it is easy to release the effect of different blockage ratios. The upstream water level rises by increasing the blockage ratio. Increasing upstream water level may cause flooding over the banks of the waterway. In the 70% blockage case, the upstream water level rises to 2.3 times of culvert height more than the non-blocked case at the same discharge and submerged ratio. The water surface profile shows an increase in water level upstream the culvert due to a decrease in transverse velocity. Because of decreasing velocity downstream culvert, there is an increase in water level before it reaches its uniform depth.
7.3. Velocity vectors
Scour downstream hydraulic structures mainly affects by velocities distribution and bed shear stress. Fig. 11 shows the velocity vectors and their magnitude in xz plane at the same flow conditions. The difference in the upstream water level due to the different blockage ratios is so clear. The maximum water level is in B70 and the minimum level is in B0. The inlet mean velocity value is about 0.88 m/s in B0 increases to 2.86 m/s in B70. As the blockage ratio increases, the inlet velocity increases. The outlet velocity in B0 case makes downward jet causes scour hole just after the fixed apron in the middle of the bed while the blockage causes upward water flow that appears clearly in B70. The upward jet decreases the scour depth to 0.13 culvert height less than B0 case. After the scour hole, the velocity decreases and the flow becomes uniform.
7.4. Velocity distribution
Fig. 12 represents flow velocity (Vx) distribution along the vertical depth (z/hu) upstream the inlet for the different blockage ratios at the same flow conditions. From the Figure, the maximum velocity creates closed to bed in B0 while in blocked case, the maximum horizontal velocity creates at 0.30 of relative vertical depth (z/hu). Fig. 13 shows the (Vz) distribution along the vertical depth (z/hu) upstream culvert at sec 1. From the mentioned Figure, it is easy to note that the maximum vertical is in B70 which appears that as the blockage ratio increases the vertical ratio also increases. In the non-blocked case. The vertical velocity (Vz) is maximum at (z/hu) equals 0.64. At the end of the fixed apron (sec 3), the horizontal velocity (Vx) is slowly increasing to reach the maximum value closed to bed in B0 and B30 while the maximum horizontal velocity occurs near to the top surface in B50 and B70 as shown in Fig. 14. The vertical velocity component along the vertical depth (z/hd) is presented in Fig. 15. The vertical velocity (Vz) is maximum in B0 at vertical depth (z/hd) 0.3 with value 0.45 m/s downward. Figs. 16 and 17 observe velocity components (Vx, Vz) along the vertical depth just after the end of blockage length at the centerline of the culvert barrel. It could be noticed the uniform velocity distribution in B0 case with horizontal velocity (Vx) closed to 1.0 m/s and vertical velocity closed to zero. In the blocked case, the maximum horizontal velocity occurs in depth more than the blockage height.
7.5. Bed velocity distribution
Fig. 18 presents the x-velocity vectors at 1.5 cm above the bed for different blockage ratios from the velocity vectors distribution and magnitude, it is easy to realize the position of the scour hole and deposition region. In B0 and B30, the flow is symmetric so that the scour hole is created around the centerline of flow while in B50 and B70 cases, the flow is asymmetric and the scour hole creates in the right of flow direction in B50. The maximum scour depth is found in the left of flow direction in B70 case where the high velocity region is found.
8. Maximum scour depth prediction
Regression analysis is used to estimate maximum scour depth downstream box culvert for different ratios of blockage by correlating the maximum relative scour by other variables that affect on it in one formula. An equation is developed to predict maximum scour depth for blocked and non-blocked. As shown in the equation below, the relative maximum scour depth(ds/hd) is a function of densimetric Froude number (Fd), blockage ratio (B) and submerged ratio (S)(11)dsh=0.56Fd-0.20B+0.45S-1.05
In this equation the coefficient of correlation (R2) is 0.82 with standard error equals 0·08. The developed equation is valid for Fd = [0.9 to 2.10] and submerged ratio (S) ≥ 1.00. Fig. 19 shows the comparison between relative maximum scour depths (ds/h) measured and estimated for different blockage ratios. Fig. 20 clears the comparison between residuals and ds/h estimated for the present study. From these figures, it could be noticed that there is a good agreement between the measured and estimated relative scour depth.
9. Comparison with previous scour equations
Many previous scour formulae have been produced for calculation the maximum scour depth downstream non-blockage culvert. These equations have been included the effect of flow regime, culvert shape, soil properties and the flow rate on maximum scour depth. Two of previous experimental studies data have been chosen to be compared with the present study results in non-blocked study data. Table 2 shows comparison of culvert shape, densmetric Froude number, median particle size and scour equations for these previous studies. By applying the present study data in these studies scour formula as shown in Fig. 21, it could be noticed that there are a good agreement between present formula results and others empirical equations results. Where that Lim [44] and Abt [4] are so closed to the present study data.
Table 2. Comparison of some previous scour formula.
The present study has shown that the FLOW 3D model can accurately simulate water surface and the scour hole characteristics downstream the box culvert with error percentage in water depths does not exceed 2.37%. Velocities distribution through and outlets culvert barrel helped on understanding the scour hole shape.
The blockage through culvert had caused of increasing of water surface upstream structure where the upstream water level in B70 was 2.3 of culvert height more than non-blocked case at the same discharge that could be dangerous on the stability of roads above. The depth averaged velocity through culvert barrel increased by 3 times its value in non-blocked case.
On the other hand, blockage through culvert had a limited effect on the maximum scour depth. The little effect of blockage on maximum scour depth could be noticed in Fig. 11. From this Figure, it could be noted that the residual part of culvert barrel after the blockage part had made turbulences. These turbulences caused the deviation of the flow resulting in the formation of asymmetric scour hole on the side of channel. This not only but in B70 the blockage height caused upward jet which made a wide far scour hole as cleared from the results in Table 1.
An empirical equation was developed from the results to estimate the maximum scour depth relative to culvert height function of blockage ratio (B), submerged ratio (S), and densimetric Froude number (Fd). The equation results was compared with some scour formulas at the same densimetric Froude number rang where the present study results was in between the other equations results as shown in Fig. 21.
Declaration of Competing Interest
The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.
References
[1]P. Sarathi, M. Faruque, R. BalachandarInfluence of tailwater depth, sediment size and densimetric Froude number on scour by submerged square wall jetsJ. Hydraul. Res., 46 (2) (2008), pp. 158-175CrossRefView Record in ScopusGoogle Scholar[2]H. Abida, R. TownsendLocal scour downstream of box-culvert outletsJ. Irrig. Drain. Eng., 117 (3) (1991), pp. 425-440CrossRefView Record in ScopusGoogle Scholar[3]S.R. Abt, C.A. Donnell, J.F. Ruff, F.K. DoehringCulvert Slope and Shape Effects on Outlet ScourTransp. Res. Rec., 1017 (1985), pp. 24-30View Record in ScopusGoogle Scholar[4]S.R. Abt, R.L. Kloberdanz, C. MendozaUnified culvert scour determinationJ. Hydraul. Eng., 110 (10) (1984), pp. 1475-1479CrossRefView Record in ScopusGoogle Scholar[5]J.P. Bohan, Erosion And Riprap Requirements At Culvert And Storm-Drain Outlets, ARMY ENGINEER WATERWAYS EXPERIMENT STATION VICKSBURG MISS1970.Google Scholar[6]C. Mendoza, S.R. Abt, J.F. RuffHeadwall influence on scour at culvert outletsJ. Hydraul. Eng., 109 (7) (1983), pp. 1056-1060CrossRefView Record in ScopusGoogle Scholar[7]H. Breusers, A. Raudkivi, Scouring, hydraulic structures design manual, vol. 143, IAHR, AA Balkema, Rotterdam, 1991.Google Scholar[8]K. Ali, S. LimLocal scour caused by submerged wall jetsProc. Inst. Civ. Eng., 81 (4) (1986), pp. 607-645CrossRefView Record in ScopusGoogle Scholar[9]O. Aderibigbe, N. RajaratnamEffect of sediment gradation on erosion by plane turbulent wall jetsJ. Hydraul. Eng., 124 (10) (1998), pp. 1034-1042View Record in ScopusGoogle Scholar[10]F.W. Blaisdell, C.L. AndersonA comprehensive generalized study of scour at cantilevered pipe outletsJ. Hydraul. Res., 26 (4) (1988), pp. 357-376CrossRefView Record in ScopusGoogle Scholar[11]Y.-M. Chiew, S.-Y. LimLocal scour by a deeply submerged horizontal circular jetJ. Hydraul. Eng., 122 (9) (1996), pp. 529-532View Record in ScopusGoogle Scholar[12]R.A. Day, S.L. Liriano, W.R. WhiteEffect of tailwater depth and model scale on scour at culvert outletsProc. Instit. Civil Eng. – Water Marit. Eng., 148 (3) (2001), pp. 189-198http://www.icevirtuallibrary.com/doi/10.1680/wame.2001.148.3.189, 10.1680/wame.2001.148.3.189View Record in ScopusGoogle Scholar[13]S. Emami, A.J. SchleissPrediction of localized scour hole on natural mobile bed at culvert outletsScour and Erosion (2010), pp. 844-853CrossRefView Record in ScopusGoogle Scholar[14]S. Sorourian, A. Keshavarzi, J. Ball, B. SamaliStudy of Blockage Effect on Scouring Pattern Downstream of a Box Culvert under Unsteady FlowAustr. J Water Resor. (2013)Google Scholar[15]S. Sorourian, Turbulent Flow Characteristics At The Outlet Of Partially Blocked Box Culverts, in: 36th IAHR World Congress, The Hague, the Netherlands, 2015.Google Scholar[16]J. Ruff, S. Abt, C. Mendoza, A. Shaikh, R. KloberdanzScour at culvert outlets in mixed bed materialsUnited States. Federal Highway Administration. Office of Research and Development (1982)Google Scholar[17]S.A. Ansari, U.C. Kothyari, K.G.R. RajuInfluence of cohesion on scour under submerged circular vertical jetsJ. Hydraul. Eng., 129 (12) (2003), pp. 1014-1019View Record in ScopusGoogle Scholar[18]B. Crookston B. Tullis, Scour and Riprap Protection in a Bottomless Arch Culvert, in: World Environmental and Water Resources Congress 2008: Ahupua’A, 2008, pp. 1–10.Google Scholar[19]S.R. Abt, J. Ruff, F. Doehring, C. DonnellInfluence of culvert shape on outlet scourJ. Hydraul. Eng., 113 (3) (1987), pp. 393-400View Record in ScopusGoogle Scholar[20]Y.H. Chen, Scour at outlets of box culverts, Colorado State University, 1970.Google Scholar[21]S. Abt, P. Thompson, T. LewisEnhancement of the culvert outlet scour estimation equationsTransp. Res. Rec. J. Transp. Res. Board, 1523 (1996), pp. 178-185View Record in ScopusGoogle Scholar[22]F.K. Doehring, S.R. AbtDrop height influence on outlet scourJ. Hydraul. Eng., 120 (12) (1994), pp. 1470-1476CrossRefView Record in ScopusGoogle Scholar[23]W. Weeks, A. Barthelmess, E. Rigby, G. Witheridge, R. Adamson, Australian rainfall and runoff revison project 11: blockage of hydraulic structures, 2009.Google Scholar[24]W. Weeks, G. Witheridge, E. Rigby, A. BarthelmessProject 11: blockage of hydraulic structuresEngineers Australia (2013)Google Scholar[25]S.R. Abt, T.E. Brisbane, D.M. Frick, C.A. McKnightTrash rack blockage in supercritical flowJ. Hydraul. Eng., 118 (12) (1992), pp. 1692-1696View Record in ScopusGoogle Scholar[26]E. Rigby, M. Boyd, S. Roso, P. Silveri, A. Davis, Causes and effects of culvert blockage during large storms, in: Global solutions for urban drainage, 2002, pp. 1–16.Google Scholar[27]S. Roso, M. Boyd, E. Rigby, R. VanDrie“Prediction of increased flooding in urban catchments due to debris blockage and flow diversionsProceedings Novatech (2004), pp. 8-13View Record in ScopusGoogle Scholar[28]C.-D. Jan, C.-L. ChenDebris flows caused by Typhoon Herb in Taiwanin Debris-Flow Hazards and Related Phenomena, Springer (2005), pp. 539-563CrossRefGoogle Scholar[29]L.W. Zevenbergen, P.F. Lagasse, P.E. Clopper, Effects of debris on bridge pier scour, in: World Environmental and Water Resources Congress 2007: Restoring Our Natural Habitat, 2007, pp. 1–10.Google Scholar[30]A. Barthelmess, E. Rigby, Estimating Culvert and Bridge Blockages-a Simplified Procedure, in: Proceedings of the 34th World Congress of the International Association for Hydro-Environment Research and Engineering: 33rd Hydrology and Water Resources Symposium and 10th Conference on Hydraulics in Water Engineering, Engineers Australia, 2011, pp. 39.Google Scholar[31]E. Rigby, A. Barthelmess, Culvert Blockage Mechanisms and their Impact on Flood Behaviour, in: Proceedings of the 34th World Congress of the International Association for Hydro-Environment Research and Engineering: 33rd Hydrology and Water Resources Symposium and 10th Conference on Hydraulics in Water Engineering, Engineers Australia, 2011, pp. 380.Google Scholar[32]G. Streftaris, N. Wallerstein, G. Gibson, S. ArthurModeling probability of blockage at culvert trash screens using Bayesian approachJ. Hydraul. Eng., 139 (7) (2012), pp. 716-726Google Scholar[33]R. Jaeger, T. LuckeInvestigating the relationship between rainfall intensity, catchment vegetation and debris mobilityInt. J. GEOMATE, 12 (33) (2017), pp. 22-29 Download PDFView Record in ScopusGoogle Scholar[34]S. Amiraslani, J. Fahimi, H. Mehdinezhad, The Numerical Investigation of Free Falling Jet’s Effect On the Scour of Plunge Pool, in: XVIII International conference on water resources, Tehran University, Iran, 2008.Google Scholar[35]A.W. Nielsen, X. Liu, B.M. Sumer, J. FredsøeFlow and bed shear stresses in scour protections around a pile in a currentCoast. Eng., 72 (2013), pp. 20-38ArticleDownload PDFView Record in ScopusGoogle Scholar[36]G. Epely-Chauvin, G. De Cesare, S. SchwindtNumerical modelling of plunge pool scour evolution in non-cohesive sedimentsEng. Appl. Comput. Fluid Mech., 8 (4) (2014), pp. 477-487 Download PDFCrossRefView Record in ScopusGoogle Scholar[37]H. Karami, H. Basser, A. Ardeshir, S.H. HosseiniVerification of numerical study of scour around spur dikes using experimental dataWater Environ. J., 28 (1) (2014), pp. 124-134CrossRefView Record in ScopusGoogle Scholar[38]S.-H. Oh, K.S. Lee, W.-M. JeongThree-dimensional experiment and numerical simulation of the discharge performance of sluice passageway for tidal power plantRenew. Energy, 92 (2016), pp. 462-473ArticleDownload PDFView Record in ScopusGoogle Scholar[39]M.A. Khodier, B.P. TullisExperimental and computational comparison of baffled-culvert hydrodynamics for fish passageJ. Appl. Water Eng. Res. (2017), pp. 1-9CrossRefView Record in ScopusGoogle Scholar[40]F.S. Inc., FLOW-3D user’s manual, Flow Science, Inc., 2009.Google Scholar[41]G. Wei, J. Brethour, M. Grünzner, J. BurnhamSedimentation scour modelFlow Science Report, 7 (2014), pp. 1-29View Record in ScopusGoogle Scholar[42]R. Soulsby, R. Whitehouse, Threshold of sediment motion in coastal environments, in: Pacific Coasts and Ports’ 97: Proceedings of the 13th Australasian Coastal and Ocean Engineering Conference and the 6th Australasian Port and Harbour Conference, vol. 1, Centre for Advanced Engineering, University of Canterbury, 1997, pp. 145.Google Scholar[43]L.C.v. RijnSediment transport, part II: suspended load transportJ. Hydraul. Eng., 110 (11) (1984), pp. 1613-1641View Record in ScopusGoogle Scholar[44]S Y LIMScour below unsubmerged full-flowing culvert outletsProc. Instit. Civil Eng. – Water Marit. Energy, 112 (2) (1995), pp. 136-149http://www.icevirtuallibrary.com/doi/10.1680/iwtme.1995.27659, 10.1680/iwtme.1995.27659View Record in ScopusGoogle Scholar
Peer review under responsibility of Faculty of Engineering, Alexandria University.
Rasoul Daneshfaraz*, Ehsan Aminvash**, Silvia Di Francesco***, Amir Najibi**, John Abraham****
토목공학의 수치해석법
Abstract
The main purpose of this study is to provide a method to increase energy dissipation on an inclined drop. Therefore, three types of rough elements with cylindrical, triangular and batshaped geometries are used on the inclined slope in the relative critical depth range of 0.128 to 0.36 and the effect of the geometry of these elements is examined using Flow 3D software. The results showed demonstrate that the downstream relative depth obtained from the numerical analysis is in good agreement with the laboratory results. The application of rough elements on the inclined drop increased the downstream relative depth and also the relative energy dissipation. The application of rough elements on the sloping surface of the drop significantly reduced the downstream Froude number, so that the Froude number in all models ranging from 4.7~7.5 to 1.45~3.36 also decreased compared to the plain drop. Bat-shaped elements are structurally smaller in size, so the use of these elements, in addition to dissipating more energy, is also economically viable.
이 연구의 주요 목적은 경사진 낙하에서 에너지 소산을 증가시키는 방법을 제공하는 것입니다. 따라서 0.128 ~ 0.36의 상대 임계 깊이 범위에서 경사면에 원통형, 삼각형 및 박쥐 모양의 형상을 가진 세 가지 유형의 거친 요소가 사용되며 이러한 요소의 형상의 영향은 Flow 3D 소프트웨어를 사용하여 조사됩니다. 결과는 수치 분석에서 얻은 하류 상대 깊이가 실험실 결과와 잘 일치함을 보여줍니다. 경 사진 낙하에 거친 요소를 적용하면 하류 상대 깊이와 상대 에너지 소산이 증가했습니다. 낙차 경사면에 거친 요소를 적용하면 하류의 Froude 수를 크게 감소시켜 4.7~7.5에서 1.45~3.36 범위의 모든 모델에서 Froude 수도 일반 낙차에 비해 감소했습니다. 박쥐 모양의 요소는 구조적으로 크기가 더 작기 때문에 더 많은 에너지를 분산시키는 것 외에도 이러한 요소를 사용하는 것이 경제적으로도 가능합니다.
Keywords: Downstream depth, Energy dissipation, Froude number, Inclined drop, Roughness elements
Introduction
급수 네트워크 시스템, 침식 수로, 수처리 시스템 및 경사가 큰 경우 흐름 에너지를 더 잘 제어하기 위해 경사 방울을 사용할 수 있습니다. 낙하 구조는 지반의 자연 경사를 설계 경사로 변환하여 에너지 소산, 유속 감소 및 수심 증가를 유발합니다. 따라서 흐름의 하류 에너지를 분산 시키기 위해 에너지 분산 구조를 사용할 수 있습니다. 난기류와 혼합된 물과 공기의 형성은 에너지 소비를 증가 시키는 효과적인 방법입니다. 흐름 경로에서 거칠기 요소를 사용하는 것은 에너지 소산을 위한 알려진 방법입니다. 이러한 요소는 흐름 경로에 배치됩니다. 그들은 종종 에너지 소산을 증가시키기 위해 다른 기하학적 구조와 배열을 가지고 있습니다. 이 연구의 목적은 직사각형 경사 방울에 대한 거칠기 요소의 영향을 조사하는 것입니다.
Conclusions
현재 연구에서 FLOW-3D 소프트웨어를 사용하여 한 높이, 한 각도, 밀도 15% 및 지그재그 배열에서 삼각형, 원통형 및 박쥐 모양의 형상을 가진 세 가지 유형의 거칠기 요소를 사용하여 경사 낙하 수리학적 매개변수에 대한 거칠기 요소 형상의 영향 평가되었다. VOF 방법을 사용하여 자유 표면 흐름을 시뮬레이션하고 초기에 3개의 난류 모델 RNG, k-ɛ 및 kω를 검증에 사용하고 이를 검토한 후 RNG 방법을 사용하여 다른 모델을 시뮬레이션했습니다. 1- 수치 결과에서 얻은 부드러운 경사 방울의 하류 상대 깊이는 실험실 데이터와 매우 좋은 상관 관계가 있으며 원통형 요소가 장착 된 경사 방울의 상대 에지 깊이 값이 가장 높았습니다. 2- 하류 상대깊이는 임계상대깊이가 증가함에 따라 상승하는 경향을 나타내어 박쥐형 요소를 구비한 경사낙하와 완만한 경사낙하가 각각 하류상대깊이가 가장 높고 가장 낮았다. 3- 하류 깊이의 증가로 인해 상대적 임계 깊이가 증가함에 따라 상대적 에너지 소산이 감소합니다. 한편, 가장 높은 에너지 소산은 박쥐 모양의 요소가 장착된 경사 낙하와 관련이 있으며 가장 낮은 에너지 소산은 부드러운 낙하와 관련이 있습니다. 삼각형, 원통형 및 박쥐 모양의 거친 요소가 장착된 드롭은 부드러운 드롭보다 각각 65%, 76% 및 85% 더 많은 흐름 에너지를 소산합니다. 4- 낙차의 경사면에 거친 요소를 적용하여 다운 스트림 Froude 수를 크게 줄여 4.7 ~ 7.5에서 1.45 ~ 3.36까지의 모든 모델에서 Froude 수가 부드러운 낙하에 비해 감소했습니다. 또한, 다른 원소보다 부피가 작은 박쥐 모양의 거칠기의 부피로 인해 이러한 유형의 거칠기를 사용하는 것이 경제적입니다.
References
References: [1] Abbaspour, A., Shiravani, P., and Hosseinzadeh dalir, A., “Experimental study of the energy dissipation on the rough ramps”, ISH journal of hydraulic engineering, 2019, p. 1-9. [2] Abraham, J.P., Sparrow, E.M., Gorman, J.M., Zhao, Y., and Minkowycz, W.J., “Application of an Intermittency model for laminar, transitional, and turbulent internal flows”, Journal of Fluids Engineering, vol. 141, 2019, paper no. 071204. [3] Ahmad, Z., Petappa, N.M., and Westrich, B., “Energy dissipation on block ramps with staggered boulders”, Journal of hydraulic engineering, vol. 135(6), 2009, p. 522-526. [4] Babaali, H.R., Shamsai, A., and Vosoughifar, H.R., “Computational modeling of the hydraulic jump in the stilling basin with convergence walls using CFD codes”, Arabian Journal for Science and Engineering, vol. 40(2), 2014, p. 381-395. [5] Castillo, L.G., Carrillo, J.M., and Cacía, J.T., “Numerical simulations and laboratory measurements in hydraulic jumps”, International conference on hydroinformatics. (2014, August) New York city. [6] Daneshfaraz, R., Aminvash, E., Esmaeli, R., Sadeghfam, S., and Abraham, J., “Experimental and numerical investigation for energy dissipation of supercritical flow in sudden contractions”, Journal of groundwater science and engineering, vol. 8(4), 2020a, p. 396-406. [7] Daneshfaraz, R., Aminvash, E., Ghaderi, A., Kuriqi, A., and Abraham, J., “Three-dimensional investigation of hydraulic properties of vertical drop in the presence of step and grid dissipators”, Symmetry, vol. 13 (5), 2021a, p. 895. [8] Daneshfaraz, R., Aminvash, E., Ghaderi, A., Abraham, J., and Bagherzadeh, M., “SVM performance for predicting the effect of horizontal screen diameters on the hydraulic parameters of a vertical drop”, Applied sciences, vol. 11 (9), 2021b, p. 4238. [9] Daneshfaraz, R., Bagherzadeh, M., Esmaeeli, R., Norouzi, R., and Abraham, J. “Study of the performance of support vector machine for predicting vertical drop hydraulic parameters in the presence of dual horizontal screens”, Water supply, vol 21(1), 2021c, p. 217-231. [10] Daneshfaraz, R., and Ghaderi, A., “Numerical investigation of inverse curvature ogee spillways”, Civil engineering journal, vol. 3(11), 2017, p. 1146-1156. [11] Daneshfaraz, R., Majedi Asl, M., and Bagherzadeh, M., “Experimental Investigation of the Energy Dissipation and the Downstream Relative Depth of Pool in the Sloped Gabion Drop and the Sloped simple Drop”, AUT Journal of Civil Engineering, 2020b (In persian). [12] Daneshfaraz, R., Majedi Asl, M., Bazyar, A., Abraham, J., Norouzi, R., “The laboratory study of energy dissipation in inclined drops equipped with a screen”, Journal of Applied Water Engineering and Research, 2020c, p. 1-10. [13] Daneshfaraz, R., Minaei, O., Abraham, J., Dadashi, S., and Ghaderi, A., “3-D Numerical simulation of water flow over a broad-crested weir with openings”, ISH Journal of Hydraulic Engineering, 2019, p.1-9. [14] Daneshfaraz, R., Sadeghfam, S., and Kashani, M., “Numerical simulation of flow over stepped spillways”, Research in civil engineering and environmental engineering, vol. 2(4), 2014, p. 190-198. [15] Ghaderi, A., Abbasi, S., Abraham, J., and Azamathulla, H.M., “Efficiency of trapezoidal labyrinth shaped stepped spillways”, Flow measurement and instrumentation, vol. 72, 2020a. [16] Ghaderi, A., Daneshfaraz, R., Dasineh, M., and Di Francesco, S., “Energy dissipation and hydraulics of flow over trapezoidaltriangular labyrinth weirs”, Water, vol. 12(7), 2020b, p. 1-18. [17] Ghaderi, A., Daneshfaraz, R., Torabi, M., Abraham, and Azamathulla, H.M. “Experimental investigation on effective scouring parameters downstream from stepped spillways”, Water supply, vol. 20(4), 2020c, p. 1-11. [18] Ghare, A.D., Ingle, R.N., Porey, P.D., and Gokhale, S.S. “Block ramp design for efficient energy dissipation”, Journal of energy dissipation, vol. 136(1), 2010, p. 1-5. [19] Gorman, J.M., Sparrow, E.M., Smith, C.J., Ghoash, A., Abraham, J.P., Daneshfaraz, R., Rezezadeh, J., “In-bend pressure drop and post-bend heat transfer for a bend with partial blockage at its inlet”, Numerical Heat Transfer A, vol, 73, 2018, p. 743-767. [20] Jamil, M., and Khan, S.A., “Theorical study of hydraulic jump in circular channel section”, ISH journal of hydraulic engineering, vol. 16(1), 2010, p. 1-10. [21] Katourani, S., and Kashefipour, S.M., “Effect of the geometric characteristics of baffle on hydraulic flow condition in baffled apron drop”, Irrigation sciences and engineering, vol. 37(2), 2012, p. 51-59. [22] Lai, Y.G., and Wu, K.A., “Three-dimensional flow and sediment transport model for free surface open channel flow on unstructured flexible meshes”, Fluids, vol. 4(1), 2019, p. 1-19.
[23] Nayebzadeh, B., Lotfollahi yaghin, M.A., and Daneshfaraz, R., “Numerical investigation of hydraulic characteristics of vertical drops with screens and gradually wall expanding”, Amirkabir journal of civil engineering, 2020 (In Persian). [24] Nurouzi, R., Daneshfaraz, R., and Bazyar, A., “The study of energy dissipation due to the use of vertical screen in the downstream of inclined drop by adaptive Neuro-Fuzzy inference system (ANFIS)”, AUT journal of civil engineering, 2019, (In Persian). [25] Ohtsu, I., and Yasuda, Y., “Hydraulic jump in sloping channel”, Journal of hydraulic engineering, vol. 117(7), 1991, p. 905-921. [26] Olsen, L., Abraham, J.P., Cheng, L.K., Gorman, J.M., and Sparrow, E.M., “Summary of forced-convection fluid flow and heat transfer for square cylinders of different aspect ratios ranging from the cube to a two-dimensional cylinder”, Advances in Heat Transfer, Vol. 51, 2019, p. 351-457. [27] Pagliara, S., Das, R., and Palermo, M., “Energy dissipation on submerged block ramps”, Journal of irrigation and drainage engineering, vol. 134(4), 2008, p.527-532. [28] Pagliara, S., and Palermo, M., “Effect of stilling basin geometry on the dissipative process in the presence of block ramps”, Journal of irrigation and drainage engineering, vol. 138(11), 2012, p. 1027-1031. [29] Simsek, O., Akoz, M.S, and Soydan, N.G., “Numerical validation of open channel flow over a curvilinear broad-creasted weir”, Progress in computational fluid dynamics an international journal, vol. 16(6), 2016, p. 364-378. [30] Sharif, N., and Rostami, A., “Experimental and numerical study of the effect of flow sepration on dissipating energy in compound bucket”, APCBEE procedia, vol. 9, 2014, p. 334-338. [31] Sparrow, E.M., Tong, J.C.K., and Abraham, J.P., “Fluid flow in a system with separate laminar and turbulent zones”, Numerical Heat Transfer A, vol. 53(4), 2008, p. 341-353. [32] Sparrow, E.M., Gorman, J.M., Abraham, J.P., and Minkowycz, W.J., “Validation of turbulence models for numerical simulation of fluid flow and convective heat transfers”, Advances in Heat Transfer, vol. 49, 2017, p. 397-421. [33] Wagner, W.E., “Hydraulic model studies of the check intake structure-potholes East canal”, Bureau of reclamation hydraulic laboratory report hyd, 1956, 411.
The most widely used method of flushing of reservoirs is to remove the deposited sediment through the bottom outlets. The size and shape of gates affect the outflow volume of water, the volume of removed sediments, and flushing efficiency. The purpose of this study is to investigate the effect of the area, number and shape of the bottom outlet gates on the velocity, concentration, and volume of the removed sediments and the dimensions of the flushing cone. Four different shapes with the same area were used for this purpose. Moreover, to study the effect of area and number of gates on flushing efficiency, circular gates with two different diameters were used. In this research, various pressure flushing modes were simulated using the Flow-3D model. Calibration and evaluation of this model were performed based on experimental findings. Results showed the parameters of the Flow-3D measures such as length, width, maximum depth, and flushing cone size with an average error of 3%, which is in good agreement with experimental results. As the area of the outlet gates increases, flushing is less risky in viewpoints of the operation process. Furthermore, the gate with a horizontal-rectangular section has an optimal shape with the highest flushing efficiency.
저수지를 세척하는 가장 널리 사용되는 방법은 바닥 배출구를 통해 침전된 침전물을 제거하는 것입니다. 게이트의 크기와 모양은 물의 유출량, 제거 된 퇴적물의 양 및 세척 효율에 영향을 미칩니다.
이 연구의 목적은 제거된 퇴적물의 속도, 농도 및 부피와 플러싱 콘의 크기에 대한 바닥 출구 게이트의 면적, 수 및 모양의 영향을 조사하는 것입니다.
이 목적을 위해 동일한 면적을 가진 4 개의 다른 모양이 사용되었습니다. 또한 플러싱 효율에 대한 면적과 게이트 수의 영향을 연구하기 위해 두 가지 직경의 원형 게이트를 사용했습니다. 이 연구에서는 Flow-3D 모델을 사용하여 다양한 압력 플러싱 모드를 시뮬레이션했습니다.
이 모델의 보정 및 평가는 실험 결과를 기반으로 수행되었습니다. 결과는 길이, 너비, 최대 깊이 및 플러싱 콘 크기와 같은 Flow-3D 측정의 매개 변수를 보여 주며 평균 오차는 3 %로 실험 결과와 잘 일치합니다. 출구 게이트의 면적이 증가함에 따라 작동 과정의 관점에서 플러싱이 덜 위험합니다. 또한 수평 직사각형 단면의 게이트는 최고의 세척 효율로 최적의 모양을 갖습니다.
Keywords
Computer model
Scouring
Flushing
Bottom outlet
Flow-3D
Sedimentation
References
Atkinson E (1996) The feasibility of flushing sediment from the reservoir. Report OD 137. Wallingford.
Brandt SA (2000) A review of reservoir desiltation. International Journal of Sediment Research. 15:321–342Google Scholar
Brethour J, Burnham J (2010) Modeling sediment erosion and deposition with the FLOW-3D sedimentation & scour model. Flow Science Technical Note. FSI-10-TN85, pp: 1-22.
Dawdy DR, Vanoni VA (1986) Modeling alluvial channels. Water Resources Research. Vol. 22(9S):71S–81SGoogle Scholar
Dehghani AA, Mosaedi A, Imamgholizadeh S, Meshkati ME (2010) Experimental investigation of pressure flushing technique in reservoir storages. Application Plans of the Ministry of Energy
Esmaeili T, Sumi T, Kantoush SA, Kubota Y, Haun S, Rüther N (2017) Three-dimensional numerical study of free-flow sediment flushing to increase the flushing efficiency: a case-study reservoir in Japan. Water. Vol. 9. No. 11, p. 900. https://doi.org/10.3390/w9110900 .
Fang D, Cao S (1996) An experimental study on scour funnel in front of a sediment flushing outlet of a reservoir. Proceedings of the 6th Federal Interagency Sedimentation Conference. Las Vegas. March 10-14, pp: I.78-I.84.
Hemphil RG (1931) Silting and life of southwestern reservoirs. Proceedings of the American Society of Civil Engineers. 56(5):967–980Google Scholar
Holly FM, Cunge JA (1975) Time dependent mass dispersion in natural streams. In: Modelling Techniques. ASCE, San Francisco, pp 1121–1137Google Scholar
Huan CC, Lai JS, Lee FZ, Tan Y C (2018) Physical model-based investigation of reservoir sedimentation processes. Water. Vol. 10, No. 4, p. 352. https://doi.org/10.3390/w10040352.
Lyn H (1987) Unsteady sediment transport modeling. Journal of Hydraulic Engineering. ASCE 110(4):450–466Google Scholar
Meshkati ME, Dehghani AA, Naser G, Emamgholizadeh S, Mosaedi A (2009) Evolution of developing flushing cone during the pressurized flushing in reservoir storage. World Academy of Science. Engineering and Technology 58:1107–1111Google Scholar
Morris GL (1995) Reservoir sedimentation and sustainable development in India: problem scope and remedial strategies. Sixth International Symposium on River Sedimentation, Management of Sediment: Philosophy, Aims, and Techniques, New Delhi.
Morris GL, Fan J (1998) Reservoir sedimentation handbook: design and management of dams, reservoirs, and watersheds for sustainable use. McGraw Hill, New York. USAGoogle Scholar
Sawadogo O, Basson GR, Schneiderbauer S (2019) Physical and coupled fully three-dimensional numerical modeling of pressurized bottom outlet flushing processes in reservoirs. International Journal of Sediment Research. 34:461–474. https://doi.org/10.1016/j.ijsrc.2019.02.001ArticleGoogle Scholar
Scheuerlein H, Tritthart M, Nunez-Gonzalez F (2004) Numerical and physical modeling concerning the removal of sediment deposits from reservoirs. Conference proceeding of Hydraulic of Dams and River Structures, Tehran, Iran, pp 245–254Google Scholar
Török GT, Baranya S, Rüther N (2017) 3D CFD modeling of local scouring, bed armoring and sediment deposition. Water. Vol. 9, No. 1, p. 56, https://doi.org/10.3390/w9010056.
White WR, Bettess R (1984) The feasibility of flushing sediments through reservoirs. challenges in African hydrology and water resources Proceedings of the Harare Symposium, IAHS Publication, No.144, pp. 577-587.
Xie Z (2011) Theoretical and numerical research on sediment transport in pressurized flow conditions. The University of Nebraska-Lincoln.
Yucel O, Graf WH (1973) Bed load deposition and delta formation: a mathematical model. December 1973. Fritz Laboratory Reports. 2062.
by Hui Hu,Jianfeng Zhang andTao Li * State Key Laboratory Base of Eco-Hydraulic Engineering in Arid Area, School of Water Resources and Hydropower, Xi’an University of Technology, Xi’an 710048, China *Author to whom correspondence should be addressed. Appl. Sci.2018, 8(12), 2456; https://doi.org/10.3390/app8122456Received: 14 October 2018 / Revised: 20 November 2018 / Accepted: 29 November 2018 / Published: 2 December 2018
Abstract
The objective of this study was to evaluate the applicability of a flow model with different numbers of spatial dimensions in a hydraulic features solution, with parameters such a free surface profile, water depth variations, and averaged velocity evolution in a dam-break under dry and wet bed conditions with different tailwater depths. Two similar three-dimensional (3D) hydrodynamic models (Flow-3D and MIKE 3 FM) were studied in a dam-break simulation by performing a comparison with published experimental data and the one-dimensional (1D) analytical solution. The results indicate that the Flow-3D model better captures the free surface profile of wavefronts for dry and wet beds than other methods. The MIKE 3 FM model also replicated the free surface profiles well, but it underestimated them during the initial stage under wet-bed conditions. However, it provided a better approach to the measurements over time. Measured and simulated water depth variations and velocity variations demonstrate that both of the 3D models predict the dam-break flow with a reasonable estimation and a root mean square error (RMSE) lower than 0.04, while the MIKE 3 FM had a small memory footprint and the computational time of this model was 24 times faster than that of the Flow-3D. Therefore, the MIKE 3 FM model is recommended for computations involving real-life dam-break problems in large domains, leaving the Flow-3D model for fine calculations in which knowledge of the 3D flow structure is required. The 1D analytical solution was only effective for the dam-break wave propagations along the initially dry bed, and its applicability was fairly limited.
이 연구의 목적은 자유 표면 프로파일, 수심 변화 및 건식 및 댐 파괴에서 평균 속도 변화와 같은 매개 변수를 사용하여 유압 기능 솔루션에서 서로 다른 수의 공간 치수를 가진 유동 모델의 적용 가능성을 평가하는 것이었습니다.
테일 워터 깊이가 다른 습식베드 조건. 2 개의 유사한 3 차원 (3D) 유체 역학 모델 (Flow-3D 및 MIKE 3 FM)이 게시된 실험 데이터와 1 차원 (1D) 분석 솔루션과의 비교를 수행하여 댐 브레이크 시뮬레이션에서 연구되었습니다.
결과는 FLOW-3D 모델이 다른 방법보다 건식 및 습식 베드에 대한 파면의 자유 표면 프로파일을 더 잘 포착함을 나타냅니다. MIKE 3 FM 모델도 자유 표면 프로파일을 잘 복제했지만, 습식 조건에서 초기 단계에서 과소 평가했습니다. 그러나 시간이 지남에 따라 측정에 더 나은 접근 방식을 제공했습니다.
측정 및 시뮬레이션 된 수심 변화와 속도 변화는 두 3D 모델 모두 합리적인 추정치와 0.04보다 낮은 RMSE (root mean square error)로 댐 브레이크 흐름을 예측하는 반면 MIKE 3 FM은 메모리 공간이 적고 이 모델의 계산 시간은 Flow-3D보다 24 배 더 빠릅니다.
따라서 MIKE 3 FM 모델은 대규모 도메인의 실제 댐 브레이크 문제와 관련된 계산에 권장되며 3D 흐름 구조에 대한 지식이 필요한 미세 계산을 위해 Flow-3D 모델을 남겨 둡니다. 1D 분석 솔루션은 초기 건조 층을 따라 전파되는 댐 파괴에만 효과적이었으며 그 적용 가능성은 상당히 제한적이었습니다.
1. Introduction
저수지에 저장된 물의 통제되지 않은 방류[1]로 인해 댐 붕괴와 그로 인해 하류에서 발생할 수 있는 잠재적 홍수로 인해 큰 자연 위험이 발생한다. 이러한 영향을 최대한 완화하기 위해서는 홍수[2]로 인한 위험을 관리하고 감소시키기 위해 홍수의 시간적 및 공간적 진화를 모두 포착하여 댐 붕괴 파동의 움직임을 예측하고 댐 붕괴 파동의 전파 과정 효과를 다운스트림[3]으로 예측하는 것이 중요하다.
그러나 이러한 수량을 예측하는 것은 어려운 일이며, 댐 붕괴 홍수의 움직임을 정확하게 시뮬레이션하고 유동장에 대한 유용한 정보를 제공하기 위한 적절한 모델을 선택하는 것은 그러므로 필수적인 단계[4]이다.
적절한 수학적 및 수치적 모델의 선택은 댐 붕괴 홍수 분석에서 매우 중요한 것으로 나타났다.분석적 해결책에서 행해진 댐 붕괴 흐름에 대한 연구는 100여 년 전에 시작되었다.
리터[5]는 먼저 건조한 침대 위에 1D de 생베넌트 방정식의 초기 분석 솔루션을 도출했고, 드레슬러[6,7]와 휘담[8]은 마찰저항의 영향을 받은 파동학을 연구했으며, 스토커[9]는 젖은 침대를 위한 1D 댐 붕괴 문제에 리터의 솔루션을 확장했다.
마샬과 멩데즈[10]는 고두노프가 가스 역학의 오일러 방정식을 위해 개발한 방법론[11]을 적용하여 젖은 침대 조건에서 리만 문제를 해결하기 위한 일반적인 절차를 고안했다. Toro [12]는 습식 및 건식 침대 조건을 모두 해결하기 위해 완전한 1D 정밀 리만 용해제를 실시했다.
Chanson [13]은 특성 방법을 사용하여 갑작스러운 댐 붕괴로 인한 홍수에 대한 간단한 분석 솔루션을 연구했다. 그러나 이러한 분석 솔루션은 특히 댐 붕괴 초기 단계에서 젖은 침대의 정확한 결과를 도출하지 못했다[14,15].과거 연구의 발전은 이른바 댐 붕괴 홍수 문제 해결을 위한 여러 수치 모델[16]을 제공했으며, 헥-라스, DAMBRK, MIK 11 등과 같은 1차원 모델을 댐 붕괴 홍수를 모델링하는 데 사용하였다.
[17 2차원(2D) 깊이 평균 방정식도 댐 붕괴 흐름 문제를 시뮬레이션하는 데 널리 사용되어 왔으며[18,19,20,21,22] 그 결과 천수(shallow water) 방정식(SWE)이 유체 흐름을 나타내는 데 적합하다는 것을 알 수 있다. 그러나, 경우에 따라 2D 수치해결기가 제공하는 해결책이 특히 근거리 분야에서 실험과 일관되지 않을 수 있다[23,24]. 더욱이, 1차원 및 2차원 모델은 3차원 현상에 대한 일부 세부사항을 포착하는 데 한계가 있다.
[25]. RANS(Reynolds-averageed Navier-Stok크스 방정식)에 기초한 여러 3차원(3D) 모델이 천수(shallow water) 모델의 일부 단점을 극복하기 위해 적용되었으며, 댐 붕괴 초기 단계에서의 복잡한 흐름의 실제 동작을 이해하기 위해 사용되었다 [26,27,28]장애물이나 바닥 실에 대한 파장의 충격으로 인한 튜디 댐 붕괴 흐름 [19,29] 및 근거리 영역의 난류 댐 붕괴 흐름 거동 [4] 최근 상용화된 수치 모델 중 잘 알려진 유체 방식(VOF) 기반 CFD 모델링 소프트웨어 FLOW-3D는 컴퓨터 기술의 진보에 따른 계산력 증가로 인해 불안정한 자유 표면 흐름을 분석하는 데 널리 사용되고 있다.
이 소프트웨어는 유한 차이 근사치를 사용하여 RANS 방정식에 대한 수치 해결책을 계산하며, 자유 표면을 추적하기 위해 VOF를 사용한다 [30,31]; 댐 붕괴 흐름을 모델링하는 데 성공적으로 사용되었다 [32,33].그러나, 2D 천수(shallow water) 모델을 사용하여 포착할 수 없는 공간과 시간에 걸친 댐 붕괴 흐름의 특정한 유압적 특성이 있다.
실생활 현장 척도 시뮬레이션을 위한 완전한 3D Navier-Stokes 방정식의 적용은 더 높은 계산 비용[34]을 가지고 있으며, 원하는 결과는 천수(shallow water) 모델[35]보다 더 정확한 결과를 산출하지 못할 수 있다. 따라서, 본 논문은 3D 모델의 기능과 그 계산 효율을 평가하기 위해 댐 붕괴 흐름 시뮬레이션을 위한 단순화된 3D 모델-MIKE 3 FM을 시도한다.
MIK 3 모델은 자연 용수 분지의 여러 유체 역학 시뮬레이션 조사에 적용되었다. 보치 외 연구진이 사용해 왔다. [36], 니콜라오스 및 게오르기오스 [37], 고얄과 라토드[38] 등 현장 연구에서 유체역학 시뮬레이션을 위한 것이다. 이러한 저자들의 상당한 연구에도 불구하고, MIK 3 FM을 이용한 댐 붕괴의 모델링에 관한 연구는 거의 없었다.
또한 댐 붕괴 홍수 전파 문제를 해결하기 위한 3D 천수(shallow water)과 완전한 3D RANS 모델의 성능을 비교한 연구도 아직 보고되지 않았다. 이 공백을 메우기 위해 현재 연구의 주요 목표는 댐 붕괴 흐름을 시뮬레이션하기 위한 단순화된 3D SWE, 상세 RANS 모델 및 분석 솔루션을 평가하여 댐 붕괴 문제에 대한 정확도와 적용 가능성을 평가하는 것이다.실제 댐 붕괴 문제를 해결하기 위해 유체역학 시뮬레이션을 시도하기 전에 수치 모델을 검증할 필요가 있다.
일련의 실험 벤치마크를 사용하여 수치 모델을 확인하는 것은 용인된 관행이다. 현장 데이터 확보가 어려워 최근 몇 년 동안 제한된 측정 데이터를 취득했다.
본 논문은 Ozmen-Cagatay와 Kocaman[30] 및 Khankandi 외 연구진이 제안한 두 가지 테스트 사례에 의해 제안된 검증에서 인용한 것이다. [39] 오즈멘-카가테이와 코카만[30]이 수행한 첫 번째 실험에서, 다른 미숫물 수위에 걸쳐 초기 단계 동안 댐 붕괴 홍수파가 발생했으며, 자유 지표면 프로파일의 측정치를 제공했다. Ozmen-Cagatay와 Kocaman[30]은 초기 단계에서 Flow-3D 소프트웨어가 포함된 2D SWE와 3D RANS의 숫자 솔루션에 의해 계산된 자유 표면 프로필만 비교했다.
Khankandi 등이 고안한 두 번째 실험 동안. [39], 이 실험의 측정은 홍수 전파를 시뮬레이션하고 측정된 데이터를 제공하는 것을 목적으로 하는 수치 모델을 검증하기 위해 사용되었으며, 말기 동안의 자유 표면 프로필, 수위의 시간 진화 및 속도 변화를 포함한다. Khankandi 등의 연구. [39] 주로 실험 조사에 초점을 맞추었으며, 초기 단계에서는 리터의 솔루션과의 수위만을 언급하고 있다.
경계 조건(상류 및 하류 모두 무한 채널 길이를 갖는 1D 분석 솔루션에서는 실험 결과를 리터와 비교하는 것이 타당하지 않기 때문이다(건조 be)d) 또는 스토커(웨트 베드) 솔루션은 벽의 반사가 깊이 프로파일에 영향을 미쳤을 때, 그리고 참조 [39]의 실험에 대한 수치 시뮬레이션과의 추가 비교가 불량할 때. 이 논문은 이러한 문제를 직접 겨냥하여 전체 댐 붕괴 과정에서의 자유 표면 프로필, 수심 변화 및 속도 변화에 대한 완전한 비교 연구를 제시한다.
여기서 댐 붕괴파의 수치 시뮬레이션은 초기에 건조하고 습한 직사각형 채널을 가진 유한 저장소의 순간 댐 붕괴에 대해 두 개의 3D 모델을 사용하여 개발된다.본 논문은 다음과 같이 정리되어 있다. 두 모델에 대한 통치 방정식은 숫자 체계를 설명하기 전에 먼저 도입된다.
일반적인 단순화된 시험 사례는 3D 수치 모델과 1D 분석 솔루션을 사용하여 시뮬레이션했다. 모델 결과와 이들이 실험실 실험과 비교하는 방법이 논의되고, 서로 다른 수심비에서 시간에 따른 유압 요소의 변동에 대한 시뮬레이션 결과가 결론을 도출하기 전에 제시된다.
2. Materials and Methods
2.1. Data
첫째, 수평 건조 및 습식 침상에 대한 초기 댐 붕괴 단계 동안의 자유 표면 프로필 측정은 Ozmen-Cagatay와 Kocaman에 의해 수행되었다[30]. 이 시험 동안, 매끄럽고 직사각형의 수평 채널은 그림 1에서 표시한 대로 너비 0.30m, 높이 0.30m, 길이 8.9m이었다.
채널은 채널 입구에서 4.65m 떨어진 수직 플레이트(담) 즉, 저장소의 길이 L0=4.65mL0에 의해 분리되었다., 및 다운스트림 채널 L1=4.25 mL1. m저수지는 댐의 좌측에 위치하고 처음에는 침수된 것으로 간주되었다; 저수지의 초기 상류 수심 h는0 0.25m로 일정했다.
오른쪽의 초기 수심 h1h1 건식침대의 경우 0m, 습식침대의 경우 0.025m, 0.1m이므로 수심비 α=h1/h0α으로 세 가지 상황이 있었다. 0, 0.1, 0.4의 습식침대 조건은 플룸 끝에 낮은 보를 사용함으로써 만들어졌다. 물 표면 프로필은 3개의 고속 디지털 카메라(50프레임/s)를 사용하여 초기에 관찰되었으며, 계측 측정의 정확도는 참고문헌 [30]에서 입증되었다. In the following section, the corresponding numerical results refer to positions x = −1 m (P1), −0.5 m (P2), −0.2 m (P3), +0.2 m (P4), +0.5 m (P5), +1 m (P6), +2 m (P7), and +2.85 m (P8), where the origin of the coordinate system x = 0 is at the dam site. 3수심비 ααα 0, 0.1, 0.4의 경우 x,yx의 경우 좌표는 h0.으로 정규화된다.
<중략> ……
Table 2. RMSE values for the free surface profiles observed by Khankandi et al. [39].
Table 3. RMSE values for the water depth variations observed by Khankandi et al. [39] at the late stage.
References
Gallegos, H.A.; Schubert, J.E.; Sanders, B.F. Two-dimensional high-resolution modeling of urban dam-break flooding: A case study of Baldwin Hills, California. Adv. Water Resour.2009, 32, 1323–1335. [Google Scholar] [CrossRef]
Kim, K.S. A Mesh-Free Particle Method for Simulation of Mobile-Bed Behavior Induced by Dam Break. Appl. Sci.2018, 8, 1070. [Google Scholar] [CrossRef]
Robb, D.M.; Vasquez, J.A. Numerical simulation of dam-break flows using depth-averaged hydrodynamic and three-dimensional CFD models. In Proceedings of the Canadian Society for Civil Engineering Hydrotechnical Conference, Québec, QC, Canada, 21–24 July 2015. [Google Scholar]
LaRocque, L.A.; Imran, J.; Chaudhry, M.H. 3D numerical simulation of partial breach dam-break flow using the LES and k-ε. J. Hydraul. Res.2013, 51, 145–157. [Google Scholar] [CrossRef]
Ritter, A. Die Fortpflanzung der Wasserwellen (The propagation of water waves). Z. Ver. Dtsch. Ing.1892, 36, 947–954. [Google Scholar]
Dressler, R.F. Hydraulic resistance effect upon the dam-break functions. J. Res. Nat. Bur. Stand.1952, 49, 217–225. [Google Scholar] [CrossRef]
Dressler, R.F. Comparison of theories and experiments for the hydraulic dam-break wave. Int. Assoc. Sci. Hydrol.1954, 38, 319–328. [Google Scholar]
Whitham, G.B. The effects of hydraulic resistance in the dam-break problem. Proc. R. Soc. Lond.1955, 227A, 399–407. [Google Scholar] [CrossRef]
Stoker, J.J. Water Waves: The Mathematical Theory with Applications; Wiley and Sons: New York, NY, USA, 1957; ISBN 0-471-57034-6. [Google Scholar]
Marshall, G.; Méndez, R. Computational Aspects of the Random Choice Method for Shallow Water Equations. J. Comput. Phys.1981, 39, 1–21. [Google Scholar] [CrossRef]
Godunov, S.K. Finite Difference Methods for the Computation of Discontinuous Solutions of the Equations of Fluid Dynamics. Math. Sb.1959, 47, 271–306. [Google Scholar]
Toro, E.F. Shock-Capturing Methods for Free-Surface Shallow Flows; Wiley and Sons Ltd.: New York, NY, USA, 2001. [Google Scholar]
Chanson, H. Application of the method of characteristics to the dam break wave problem. J. Hydraul. Res.2009, 47, 41–49. [Google Scholar] [CrossRef][Green Version]
Cagatay, H.; Kocaman, S. Experimental Study of Tail Water Level Effects on Dam-Break Flood Wave Propagation; 2008 Kubaba Congress Department and Travel Services: Ankara, Turkey, 2008; pp. 635–644. [Google Scholar]
Stansby, P.K.; Chegini, A.; Barnes, T.C.D. The initial stages of dam-break flow. J. Fluid Mech.1998, 374, 407–424. [Google Scholar] [CrossRef]
Soares-Frazao, S.; Zech, Y. Dam Break in Channels with 90° Bend. J. Hydraul. Eng.2002, 128, 956–968. [Google Scholar] [CrossRef]
Zolghadr, M.; Hashemi, M.R.; Zomorodian, S.M.A. Assessment of MIKE21 model in dam and dike-break simulation. IJST-Trans. Mech. Eng.2011, 35, 247–262. [Google Scholar]
Bukreev, V.I.; Gusev, A.V. Initial stage of the generation of dam-break waves. Dokl. Phys.2005, 50, 200–203. [Google Scholar] [CrossRef]
Soares-Frazao, S.; Noel, B.; Zech, Y. Experiments of dam-break flow in the presence of obstacles. Proc. River Flow2004, 2, 911–918. [Google Scholar]
Aureli, F.; Maranzoni, A.; Mignosa, P.; Ziveri, C. Dambreak flows: Acquisition of experimental data through an imaging technique and 2D numerical modelling. J. Hydraul. Eng.2008, 134, 1089–1101. [Google Scholar] [CrossRef]
Rehman, K.; Cho, Y.S. Bed Evolution under Rapidly Varying Flows by a New Method for Wave Speed Estimation. Water2016, 8, 212. [Google Scholar] [CrossRef]
Wu, G.F.; Yang, Z.H.; Zhang, K.F.; Dong, P.; Lin, Y.T. A Non-Equilibrium Sediment Transport Model for Dam Break Flow over Moveable Bed Based on Non-Uniform Rectangular Mesh. Water2018, 10, 616. [Google Scholar] [CrossRef]
Ferrari, A.; Fraccarollo, L.; Dumbser, M.; Toro, E.F.; Armanini, A. Three-dimensional flow evolution after a dam break. J. Fluid Mech.2010, 663, 456–477. [Google Scholar] [CrossRef]
Liang, D. Evaluating shallow water assumptions in dam-break flows. Proc. Inst. Civ. Eng. Water Manag.2010, 163, 227–237. [Google Scholar] [CrossRef]
Biscarini, C.; Francesco, S.D.; Manciola, P. CFD modelling approach for dam break flow studies. Hydrol. Earth Syst. Sci.2010, 14, 705–718. [Google Scholar] [CrossRef][Green Version]
Oertel, M.; Bung, D.B. Initial stage of two-dimensional dam-break waves: Laboratory versus VOF. J. Hydraul. Res.2012, 50, 89–97. [Google Scholar] [CrossRef]
Quecedo, M.; Pastor, M.; Herreros, M.I.; Merodo, J.A.F.; Zhang, Q. Comparison of two mathematical models for solving the dam break problem using the FEM method. Comput. Method Appl. Mech. Eng.2005, 194, 3984–4005. [Google Scholar] [CrossRef]
Shigematsu, T.; Liu, P.L.F.; Oda, K. Numerical modeling of the initial stages of dam-break waves. J. Hydraul. Res.2004, 42, 183–195. [Google Scholar] [CrossRef]
Soares-Frazao, S. Experiments of dam-break wave over a triangular bottom sill. J. Hydraul. Res.2007, 45, 19–26. [Google Scholar] [CrossRef]
Ozmen-Cagatay, H.; Kocaman, S. Dam-break flows during initial stage using SWE and RANS approaches. J. Hydraul. Res.2010, 48, 603–611. [Google Scholar] [CrossRef]
Vasquez, J.; Roncal, J. Testing River2D and FLOW-3D for Sudden Dam-Break Flow Simulations. In Proceedings of the Canadian Dam Association’s 2009 Annual Conference: Protecting People, Property and the Environment, Whistler, BC, Canada, 3–8 October 2009. [Google Scholar]
Ozmen-Cagatay, H.; Kocaman, S. Dam-break flow in the presence of obstacle: Experiment and CFD simulation. Eng. Appl. Comput. Fluid2011, 5, 541–552. [Google Scholar] [CrossRef]
Ozmen-Cagatay, H.; Kocaman, S.; Guzel, H. Investigation of dam-break flood waves in a dry channel with a hump. J. Hydro-Environ. Res.2014, 8, 304–315. [Google Scholar] [CrossRef]
Gu, S.L.; Zheng, S.P.; Ren, L.Q.; Xie, H.W.; Huang, Y.F.; Wei, J.H.; Shao, S.D. SWE-SPHysics Simulation of Dam Break Flows at South-Gate Gorges Reservoir. Water2017, 9, 387. [Google Scholar] [CrossRef]
Evangelista, S. Experiments and Numerical Simulations of Dike Erosion due to a Wave Impact. Water2015, 7, 5831–5848. [Google Scholar] [CrossRef][Green Version]
Bocci, M.; Chiarlo, R.; De Nat, L.; Fanelli, A.; Petersen, O.; Sorensen, J.T.; Friss-Christensen, A. Modelling of impacts from a long sea outfall outside of the Venice Lagoon (Italy). In Proceedings of the MWWD—IEMES 2006 Conference, Antalya, Turkey, 6–10 November 2006; MWWD Organization: Antalya, Turkey, 2006. [Google Scholar]
Nikolaos, T.F.; Georgios, M.H. Three-dimensional numerical simulation of wind-induced barotropic circulation in the Gulf of Patras. Ocean Eng.2010, 37, 355–364. [Google Scholar]
Goyal, R.; Rathod, P. Hydrodynamic Modelling for Salinity of Singapore Strait and Johor Strait using MIKE 3FM. In Proceedings of the 2011 2nd International Conference on Environmental Science and Development, Singapore, 26–28 February 2011. [Google Scholar]
Khankandi, A.F.; Tahershamsi, A.; Soares-Frazão, S. Experimental investigation of reservoir geometry effect on dam-break flow. J. Hydraul. Res.2012, 50, 376–387. [Google Scholar] [CrossRef]
Flow Science Inc. FLOW-3D User’s Manuals; Flow Science Inc.: Santa Fe, NM, USA, 2007. [Google Scholar]
Danish Hydraulic Institute (DHI). MIKE 3 Flow Model FM. Hydrodynamic Module-User Guide; DHI: Horsholm, Denmark, 2014. [Google Scholar]
Pilotti, M.; Tomirotti, M.; Valerio, G. Simplified Method for the Characterization of the Hydrograph following a Sudden Partial Dam Break. J. Hydraul. Eng.2010, 136, 693–704. [Google Scholar] [CrossRef]
Hooshyaripor, F.; Tahershamsi, A.; Razi, S. Dam break flood wave under different reservoir’s capacities and lengths. Sādhanā2017, 42, 1557–1569. [Google Scholar] [CrossRef]
Kocaman, S.; Ozmen-Cagatay, H. Investigation of dam-break induced shock waves impact on a vertical Wall. J. Hydrol.2015, 525, 1–12. [Google Scholar] [CrossRef]
Liu, H.; Liu, H.J.; Guo, L.H.; Lu, S.X. Experimental Study on the Dam-Break Hydrographs at the Gate Location. J. Ocean Univ. China2017, 16, 697–702. [Google Scholar] [CrossRef]
Marra, D.; Earl, T.; Ancey, C. Experimental Investigations of Dam Break Flows down an Inclined Channel. In Proceedings of the 34th World Congress of the International Association for Hydro- Environment Research and Engineering: 33rd Hydrology and Water Resources Symposium and 10th Conference on Hydraulics in Water Engineering, Brisbane, Australia, 26 June–1 July 2011. [Google Scholar]
Wang, J.; Liang, D.F.; Zhang, J.X.; Xiao, Y. Comparison between shallow water and Boussinesq models for predicting cascading dam-break flows. Nat. Hazards2016, 83, 327–343. [Google Scholar] [CrossRef]
Yang, C.; Lin, B.L.; Jiang, C.B.; Liu, Y. Predicting near-field dam-break flow and impact force using a 3D model. J. Hydraul. Res.2010, 48, 784–792. [Google Scholar] [CrossRef]
현재 scouring model은 물에잠겨있는 부분에 대해 해석을 하게되어 있으므로 packed sediment부분은 fluid region(with infinite drag)이 존재하게됩니다. 그러므로 fluid region이 없다 하더라도 packed sediment가 경사면에 존재하면 중력에 의해 내부유체의 유동이 생겨 위 예제와 같이 미소한 scouring이 표면에 물이 없는 경사면에서도 발생하는것입니다. 그러므로 이를 없애기 위해서는 물이 없는 경사면 부분은 별도의 solid로 규정하면 이 문제를 피할수 있습니다.
Tip2 ) clay가 sticky하면 일반적으로 유동의 상대운동이 감소될것이므로 drag coefficient 나 Richardson Zaki coefficient multiplier를 증가시켜 변화를 조사해 볼 수 있습니다.
FLOW-3D의 침전물 이송 모델을 사용하여 세굴 및 침전물을 평가할 수 있으며, 여기서 3차원 유량 구성 요소가 세굴 프로세스를 주도하고 있습니다. Flow-3D의 유체역학 모델은 유체물리학을 설명하는 정전기적이지 않은 레이놀즈-평균화된 Navier-Stokes 방정식을 완벽하게 해결합니다. 유체역학적 솔버는 침전물 운반 모듈과 완전히 결합되어 있어 침전물 운반 및 비접착 토양의 부유식 침식, 인포테인먼트 및 침식을 시뮬레이션합니다(Wei et al., 2014). 베드로드, 인포테인먼트 및 정착 프로세스에 사용되는 모든 경험적 관계는 완전히 사용자 정의 가능하며, 최대 10개의 침전물 종(곡물 크기, 질량 밀도, 임계 전단 응력 등 서로 다른 특성을 가진)을 정의할 수 있습니다. FLOW-3D는 짧은 경과 시간 척도에 대한 국부적 스쿠어를 시뮬레이션하는 데 이상적입니다.
FLOW-3D‘s Sediment Transport model can be used to evaluate scour and deposition, where three-dimensional flow components are driving the scouring process. FLOW-3D’s hydrodynamic model solves the full unsteady non-hydrostatic Reynolds-averaged Navier-Stokes equations that describe the flow physics. The hydrodynamic solver is fully coupled with a sediment transport module that simulates bedload and suspended sediment transport, entrainment and erosion for non-cohesive soils (Wei et al., 2014). All empirical relationships used in bedload, entrainment and settling processes are fully customizable, and up to 10 different sediment species (with different properties such as grain size, mass density and critical shear stress) can be defined. FLOW-3Dis ideal for simulating local scour over short episodic time scales.
Modeling Capabilities – Unsteady 3D mobile bed modeling – Bedload and suspended sediment transport – Non-cohesive sediment – 10 individual grain size fractions – Suspended sediment settling and entrainment – Critical angle of repose
Applications – River and coastal morphodynamics – Bridge pier and abutment scour – Local scour at hydraulic structures – Sedimentation basins – Reservoir flushing
Sediment Transport Model
Sentral Transport 모델은 8.0 버전(Brethour, 2009년)에서 처음 도입되었으며, 11.1 버전(Wei et al., 2014년), 가장 최근에는 12.0 버전(Flow Science, 2019년)에서 광범위한 개정을 거쳤습니다. 숫자 모델에서 시뮬레이션된 물리적 프로세스의 개략도가 아래에 나와 있습니다.
The Sediment Transport model was first introduced in version 8.0 (Brethour, 2009), and has gone through extensive revisions in version 11.1 (Wei et al., 2014), and most recently in version 12.0 (Flow Science, 2019). A schematic of the physical processes simulated in the numerical model is illustrated below.
수치 모델에서 침전물은 포장된 Bed로서 일시 중단된 상태로 존재할 수 있습니다. 포장된 Bed는 PRIPT™ 기법을 사용하여 복잡한 솔리드 경계(Hirt 및 Sicilian, 1985)에 표현된 지울 수 없는 솔리드 객체입니다. 이것은 유체역학 용해기의 고체 물체를 나타내는 데 사용되는 방법과 동일합니다. 포장된 Bed의 형태학적 변화는 침전물 질량의 보존에 의해 좌우됩니다.
In the numerical model, sediment can exist as packed bed and in a suspended state. A packed bed is an erodible solid object that is represented using the FAVOR™ technique for complex solid boundaries (Hirt and Sicilian, 1985). This is the same method used to represent solid objects in the hydrodynamic solver. The morphological change in the packed bed is governed by the conservation of sediment mass.
형태학적 변경은 모형에 숫자로 표시되는 여러 가지 물리적 프로세스에 의해 제어됩니다. 이러한 프로세스에는 베드로드 운송, 인포테인먼트 및 증착이 포함됩니다. 베드로드 이송은 침전물이 서스펜션에 전달되지 않고 채널을 따라 횡방향으로 이동하는 물리적 과정입니다. 인포테인먼트란 난류 에디가 패킹 베드 상단의 곡물을 제거하고 일시 중단된 상태로 전환하는 과정입니다. 포장이란 곡물이 현수막에서 안착되어 포장된 침대에 퇴적하는 과정입니다. 수치 모델에서 이것은 일시 중단된 상태에서 포장된 베드 상태로의 전환입니다.
The morphological changes are governed by several different physical processes that are represented numerically in the model. These processes include bedload transport, entrainment and deposition. Bedload transport is the physical process of sediment moving laterally along the channel without being carried into suspension. Entrainment is the process by which turbulent eddies remove the grains from the top of the packed bed and transition to the suspended state. Packing is the process of grains settling out of suspension and depositing onto the packed bed. In the numerical model, this is the transition from the suspended to the packed bed state.
인포테인먼트 및 패킹의 상대적 비율은 포장된 베드와 부유 상태 사이의 침전물 질량 교환을 제어합니다. 이 모델은 Meyer-Peter Müler(1948), Nielsen(1992) 또는 Van Rijn(1984)의 방정식을 사용하여 베드 인터페이스가 포함된 각 메시 셀에서 베드로드 전송을 계산합니다. 메쉬 셀에서 이웃의 각 메쉬 셀로 이동하는 곡물의 양을 결정하기 위해 하위 메쉬 방법이 사용됩니다. 인포테인먼트에서 곡물의 리프팅 속도는 Winterwerp 등(1992)의 방정식을 사용하여 계산됩니다. 안착 속도는 Soulsby(1997년)를 사용하여 계산됩니다. 베드 인터페이스가 포함된 메시 셀에서 인터페이스의 위치, 방향 및 면적을 계산하여 베드 전단 응력, 무차원 전단 응력, 베드로드 전송 속도 및 인포테인먼트 속도를 결정합니다. 3D 난류 흐름의 베드 전단 응력은 표준 벽 함수를 사용하여 중간 곡물 크기에 비례하는 베드 표면 거칠기를 고려하여 평가됩니다.
The relative rates of entrainment and packing control the exchange of sediment mass between the packed bed and suspended states. The model calculates bedload transport in each mesh cell containing the bed interface using the equation of Meyer-Peter Müller (1948), Nielsen (1992) or Van Rijn (1984). A sub-mesh method is employed to determine the amount of grains moving from the mesh cell into each mesh cell in its neighbor. The lifting velocity of grains in entrainment is calculated using the equation of Winterwerp et al. (1992). The settling velocity is calculated using Soulsby (1997). In the mesh cells containing the bed interface, location, orientation and area of the interface are calculated to determine the bed shear stress, dimensionless shear stress, bedload transport rates and entrainment rates. Bed shear stress in 3D turbulent flows is evaluated using the standard wall function with consideration of bed surface roughness that is proportional to the median grain size.
부유된 침전물은 유체의 스칼라 질량 농도로 표시됩니다. 농도는 주어진 셀에서 균일한 것으로 가정되며 유체 셀 밀도 및 점도와 결합됩니다. 각 종에 대해, 부유 침전물 농도는 수송 방정식을 풀어서 계산됩니다.
The suspended sediment is represented as a scalar mass concentration in the fluid. The concentration is assumed to be uniform in a given cell and is coupled with the fluid cell density and viscosity. For each species, the suspended sediment concentration is calculated by solving a transport equation.
Validations
다음 5가지 검증 사례는 실험 데이터와 FLOW-3D의 침전물 이송 모델의 시뮬레이션 결과를 비교합니다.
마오(1986년) Mao는 수중 수평 파이프라인 아래 침대의 무서운 프로파일을 얻기 위해 실험 작업을 수행했습니다. 아래 그림은 FLOW-3D를 사용하여 얻은 결과와 실험 결과를 비교합니다.
Chatterjee et al. (1994)
수평 제트 침수로 인해 국부적인 스쿠어 프로파일을 얻기 위한 실험 작업이 수행되었습니다. 아래 그림은 scour구멍 깊이와 둔부 높이에 대한 실험 대 FLOW-3D의 숫자 결과를 시간의 함수로 비교합니다. 이 애니메이션은 scour구멍과 둔부 높이가 최대 1시간 내에 안정된 상태에 도달한다는 것을 보여줍니다.
Gladstone et al. (1998)
In these experiments the propagation and deposition patterns of particle-laden flows were studied. The plot below compares experimental versus FLOW-3D simulation results from three different setups, labeled case A (100% 0.025mm size particles), case D (50% 0.069mm and 50% 0.025mm size particles), and case G (100% 0.069mm size particles).
Faruque et al. (2006)
이 논문에서, 저자들은 실험을 통해 3차원 벽면 제트기를 물에 잠기게 함으로써 국부적인 악취를 연구했습니다. 아래 표는 세 가지 서로 다른 테일워터 비율에 대한 scour 구멍의 3D 형태학적 변화에 대한 실험과 FLOW-3D 수치 결과를 비교합니다.
In this paper, the authors studied local scour by submerged three-dimensional wall jets via experiments. The table below compares the experimental versus FLOW-3D numerical results for 3D morphological changes in the scour hole for three different tailwater ratios.
References
Brethour, J.M., Hirt, C.W., 2009, Drift Model for Two-Component Flows, FSI-14-TN-83, Flow Science, Inc.
Chatterjee, S.S., Ghosh, S.N., and Chatterjee M., 1994, Local scour due to submerged horizontal jet, Journal of Hydraulic Engineering, 120(8), pp. 973-992.
Faruque, M.A.A., Sarathi, P., and Balachandar R., 2006, Clear Water Local Scour by Submerged Three-Dimensional Wall Jets : Effect of Tailwater Depth, Journal of Hydraulic Engineering, 132(6), pp. 575-580.
Flow Science, 2019, FLOW-3D Version 12.0 User Manual, Santa Fe, NM: Flow Science, Inc. https://www.flow3d.com
Gladstone, C., Phillips, J.C., and Sparks R.S.J., 1998, Experiments on bidisperse, constant-volume gravity currents: propagation and sediment deposition, Sedimentology 45, pp. 833-843.
Hirt, C.W. and Sicilian, J.M., 1985, A porosity technique for the definition of obstacles in rectangular cell meshes, 4th International Conference on Numerical Ship Hydrodynamics, Washington, D.C.
Khosronejad, A., Kang, S., & Sotiropoulos, F., 2012. Experimental and computational investigation of local scour around bridge piers, Advances in Water Resources, 37, pp. 73-85.
Mao, Y., 1986. The interaction between a pipeline and an erodible bed, PhD thesis, Institute of Hydrodynamics and Hydraulic Engineering, Technical University of Denmark, Lyngby, Denmark.
Meyer-Peter, E. and Müller, R., 1948, Formulas for bed-load transport, Proceedings of the 2nd Meeting of the International Association for Hydraulic Structures Research. pp. 39–64.
Nielsen, P., 1992, Coastal bottom boundary layers and sediment transport (Vol. 4). World scientific.
Soulsby, R., 1997, Dynamics of Marine Sands, Thomas Telford Publications, London.
Van Rijn, L. C., 1984, Sediment Transport, Part I: Bed load transport, Journal of Hydraulic Engineering 110(10), pp. 1431-1456.
Wei, G., Brethour, J.M., Grüenzner M., and Burnham, J., 2014, The Sediment Scour Model in FLOW-3D, Technical Note FSI-14-TN-99, Flow Science, Inc.
Winterwerp, J.C., Bakker, W.T., Mastbergen, D.R. and Van Rossum, H., 1992, Hyperconcentrated sand-water mixture flows over erodible bed, Journal of Hydraulic Engineering, 118(11), pp. 1508–1525.