Fact-checked by Grok 2 weeks ago

Hydraulic structure

A hydraulic structure is an engineered construction, typically submerged or partially submerged in , designed to manage, , or disrupt the natural flow of for purposes such as flood mitigation, , , and hydropower generation. These structures interact directly with hydraulic forces, including , , and , and are fundamental to in management. Hydraulic structures are classified primarily by their function, which determines their design and placement in riverine, coastal, or urban environments. Storage structures, such as and reservoirs, impound to regulate supply and generate power, with examples including large-scale embankment and gravity that can store billions of cubic meters. Diversion and regulation structures, like weirs, barrages, and , redirect or measure to canals or prevent flooding, often incorporating dissipation features such as stilling basins to avoid downstream . Conveyance structures, including aqueducts, culverts, and spillways, facilitate transport over obstacles or during high flows, with spillways specifically engineered to safely release excess from reservoirs at controlled velocities. The design of hydraulic structures emphasizes , , and environmental , often relying on physical or computational hydraulic models to predict under varying discharges and predict issues like or scour. Historical developments trace back to ancient civilizations, with early examples like the in (circa 750 BCE) demonstrating rudimentary storage and diversion techniques, evolving into modern applications that address climate challenges and sustainable water use. Today, these structures play a critical role in global , supporting on 354 million hectares through as of 2022 while mitigating flood risks in vulnerable regions.

Fundamentals

Definition and Scope

Hydraulic structures are man-made constructions, typically submerged or partially submerged in bodies of , designed to disrupt and control the natural flow of by influencing its velocity, depth, direction, storage, or conveyance. These structures are essential in for managing , mitigating flood risks, and supporting , , and needs. The scope of hydraulic structures spans multiple disciplines within , including for water quantity assessment, for flow dynamics analysis, and for foundation stability in water-saturated environments. Unlike natural water features such as or , which evolve through geological processes without human intervention, hydraulic structures are purposefully engineered to alter hydraulic regimes and integrate with broader water management systems. Key concepts in hydraulic structures revolve around modifying water behavior, such as impounding water to create reservoirs for storage or diverting flows to prevent erosion and enable utilization. Fundamental terminology includes , which represents the total energy per unit weight of water available for flow, often expressed as elevation plus ; discharge, the volume of water passing a cross-section per unit time, typically measured in cubic meters per second; and , a low barrier over which water flows to regulate or measure discharge. These structures serve primary purposes like for sustained supply during dry periods and flow diversion for agricultural or urban distribution, without overlapping into specific classifications such as or canals.

Historical Development

The earliest hydraulic structures emerged in around 3000 BCE, where basin systems harnessed the River's annual floods to support . These consisted of natural floodplains divided by earthen dikes and s into large basins, typically 9 to 106 square kilometers in size, which filled to depths of about 1.5 meters during inundation and were drained through sluice gates for crop cultivation. By the late Predynastic period (circa 3100 BCE), artificial enhancements like channel dredging and levee breaching marked the shift from purely natural to managed , as depicted in artifacts such as the Scorpion King's mace-head showing a ceremonial ditch-cutting. In the Roman era, hydraulic engineering advanced significantly with aqueducts, exemplified by the Aqua Appia constructed in 312 BCE under censor . This underground conduit, spanning 16.4 kilometers to deliver approximately 841 quinariae (about 75,000 cubic meters daily) to , represented the first major public system and set precedents for later aqueducts by integrating gravity flow and minimal above-ground exposure. During the medieval period, Islamic civilizations refined underground water conveyance through qanats—horizontal tunnels with vertical shafts—extending from aquifers to arid farmlands across regions from to , enabling sustainable and urban supply. In parallel, European water mills proliferated from the , powered by diverted streams for grinding grain and industrial tasks, with monastic orders like the optimizing wheel designs and networks for self-sufficiency. The Renaissance saw further innovation through Leonardo da Vinci's designs in around 1500, where he sketched multi-level systems like the to improve , , and mill operations by stabilizing banks and enhancing flow control. The in the 19th century introduced to dam construction, with the world's first arch dam built at 75 Miles Dam in in 1880, marking a durable alternative to masonry for water storage and . This era's engineering culminated in projects like Egypt's , completed in 1902 after construction began in 1899, which controlled floods to support perennial while generating initial . Key figures such as (1724–1792) laid foundational work in the 1760s by pioneering for watertight structures and conducting scale-model tests on waterwheels, advancing empirical hydraulic analysis. The 20th century shifted hydraulic structures toward multipurpose designs integrating flood control, power generation, and navigation, as seen in the United States' , completed in 1935 with final dedication in 1936, standing 221 meters tall and producing up to 2,074 megawatts to support regional growth during the . This trend peaked with China's , where power generation began in 2003 after reservoir filling, creating a 39.3 billion cubic meter storage for flood mitigation, 22.5 gigawatts of , and enhanced shipping for over 10,000-ton vessels. Modern hydraulic modeling pioneers, building on Smeaton's methods, further refined these projects through physical and computational simulations to optimize designs for complex river dynamics.

Classification

By Function

Hydraulic structures are classified by function according to their primary role in managing , specifically whether they modify the quantity of (e.g., by storing or diverting ), the (e.g., by controlling sediments or pollutants), or the (e.g., by conveying or regulating ). This functional emphasizes operational over material or scale, enabling engineers to select structures that address specific hydrological challenges such as supply regulation or prevention. Storage structures are engineered to impound and retain water volumes for extended periods, creating reservoirs that support seasonal or long-term water supply needs. These structures, such as dams forming conservation reservoirs, store water for irrigation, hydroelectric power generation, domestic use, or flood mitigation, ensuring availability during dry periods or peak demand. For instance, multipurpose reservoirs integrate storage for multiple uses, balancing water levels to optimize resource allocation without excessive evaporation losses. Diversion and conveyance structures redirect and transport from its natural course to desired locations, modifying the flow path to facilitate distribution. Intakes capture from or reservoirs, while flumes, aqueducts, and channels convey it over distances, often crossing obstacles like valleys or roads. Diversion weirs or barrages raise levels minimally to feed canals or pipelines, preventing significant ponding while enabling controlled abstraction for agricultural or municipal needs. Flow control structures regulate water velocity, depth, and direction to maintain stable conditions and prevent issues like flooding or deposition. Weirs and spillways manage by providing controlled release paths, reducing downstream velocities to avoid scour, while and regulators in canals adjust flows for equitable distribution. These elements ensure in systems like networks, where precise control mitigates and supports uniform water delivery. Energy dissipation structures absorb from high-velocity flows, particularly during discharges, to protect downstream areas from hydraulic forces. Stilling basins create hydraulic jumps that convert flow energy into , while chutes and baffles gradually reduce velocities, minimizing at structure toes. Such designs are critical in operations, where uncontrolled releases could otherwise destabilize riverbeds or embankments. Protection structures safeguard hydraulic systems and surrounding landscapes from floodwaters, scour, or material degradation, focusing on defensive modification of flow impacts. Levees and floodwalls contain high waters along riverbanks, while revetments—such as riprap-lined slopes—shield embankments from erosive currents. These structures enhance in vulnerable areas, like floodplains, by directing flows away from .

By Scale and Location

Hydraulic structures are classified by scale based on physical dimensions such as height and capacity metrics like discharge or storage volume, which determine their engineering demands and environmental integration. Small-scale structures typically include local installations like farm weirs or culverts with relatively modest dimensions and discharge capacities, designed for low-volume flow management in constrained settings. These are often constructed using simpler materials and techniques to handle low-volume water flows without requiring extensive foundations. In contrast, large-scale structures encompass major installations such as dams with a height of 15 meters or more from lowest foundation to crest, or between 5 and 15 meters impounding more than 3 million cubic meters of water, as defined by the International Commission on Large Dams (ICOLD) for significant water impoundment. Such structures demand advanced hydraulic modeling and robust materials to manage substantial water volumes and forces. Location further refines this classification, distinguishing between riverine and coastal placements based on hydrological regimes. Riverine structures, situated along inland waterways, include barrages that regulate river flows for diversion or , adapting to unidirectional currents and in freshwater environments. Coastal structures, conversely, such as sea walls, are positioned in tidal or estuarine zones to withstand bidirectional flows, wave action, and saline conditions, often incorporating scour protection at their bases. placements integrate structures like flood barriers into dense , prioritizing compact designs that minimize while protecting built environments from overflow. Rural settings favor agricultural diversions, such as weirs channeling water to fields, which emphasize cost-effective, low-maintenance features suited to expansive landscapes and seasonal demands. The scale and location of hydraulic structures are influenced by site-specific factors including , available water volume, and economic feasibility. Geological conditions, such as soil stability and rock formation, dictate foundation depth and , often limiting small-scale builds to favorable terrains while necessitating geotechnical reinforcements for larger ones in challenging sites. Water volume assessments, derived from hydrological data, scale structures to match expected inflows, ensuring aligns with regional patterns without excess overdesign. Economic feasibility evaluates construction costs against benefits like reliability, favoring modular small-scale options in resource-limited areas over capital-intensive large-scale projects.

Design Principles

Hydraulic Analysis

Hydraulic analysis in hydraulic structures involves applying principles of to predict water flow behavior, ensuring structures like , canals, and spillways operate safely and efficiently. This analysis builds on fundamental fluid statics, such as hydrostatic distributions, to address dynamic flow conditions where water movement influences structural performance. Key objectives include determining flow rates, velocities, water surface elevations, potential erosive effects, and risks, with the cavitation index \sigma = \frac{P - P_v}{0.5 \rho v^2} (where P is local , P_v vapor , \rho , v ) maintained above 0.2-0.5 to prevent damage, often via or profile adjustments. The foundational principles of hydraulic analysis stem from conservation laws. The , Q = A \cdot V, expresses mass conservation by relating Q (volume flow rate) to cross-sectional area A and mean V, ensuring constant flow through varying sections of a structure. Bernoulli's equation, \frac{P}{\rho g} + \frac{v^2}{2g} + z = \text{[constant](/page/Constant)}, balances total energy head—comprising , , and elevation head—along a streamline, assuming steady, incompressible, and , though losses are often incorporated for real-world applications. For open-channel flows prevalent in hydraulic structures, the , derived from Newton's second law applied to a , accounts for forces like gravity, pressure, and to describe unsteady or nonuniform , particularly useful for analyzing transitions such as discharges. Flow regimes in hydraulic structures are classified to characterize water motion and its implications for energy dissipation and wave propagation. Laminar flow occurs at low velocities where inertial forces are dominated by viscous forces, quantified by the Reynolds number Re = \frac{v D}{\nu} < 500 (with v as velocity, D as hydraulic diameter, and \nu as kinematic viscosity), resulting in smooth, layered motion rarely sustained in large-scale structures due to high flow rates. Turbulent flow, with Re > 2000, prevails in most practical scenarios, featuring chaotic eddies that enhance mixing but increase energy losses. Additionally, open-channel flows are categorized by the Froude number Fr = \frac{v}{\sqrt{g y}} (where y is flow depth and g is gravitational acceleration): subcritical flow (Fr < 1) resembles tranquil, deep-water conditions where disturbances propagate upstream; critical flow (Fr = 1) marks the transition with minimum energy for a given discharge; and supercritical flow (Fr > 1) behaves like rapid, shallow flow where waves cannot travel upstream, often occurring downstream of gates or weirs. Modeling techniques enable simulation of these principles to predict patterns without full-scale testing. One-dimensional (1D) models simplify channels as cross-sections, solving the and or equations sequentially, as implemented in software like for steady or unsteady routing in rivers and canals. Two-dimensional (2D) models extend this by incorporating lateral variations, using the to compute velocity fields and inundation over floodplains, particularly valuable for complex terrains around structures. Three-dimensional (3D) models resolve vertical variations for intricate flows, such as near piers, but demand high computational resources. Physical scale models, constructed in laboratories using Froude scaling (\lambda_v = \sqrt{\lambda_L}, where \lambda denotes scale factors for velocity and length), replicate prototype conditions to validate numerical predictions and study phenomena like scour, ensuring geometric, kinematic, and dynamic similitude. Critical parameters in hydraulic analysis quantify flow characteristics and risks. Discharge Q represents the , directly computed from the and for spillways or culverts. Head h_f = f \frac{L}{D} \frac{v^2}{2g}, via the Darcy-Weisbach (with f, length L, and diameter D), measures energy dissipation due to friction in conduits, guiding and designs in aqueducts. Scour depth estimation assesses potential around foundations, often using empirical formulas like those in HEC-18, which integrate , depth, size, and geometry to predict maximum local scour, preventing undermining during high flows.

Structural Stability

Hydraulic structures must be engineered to resist various forces imposed by and environmental conditions to ensure long-term integrity. Primary loads include hydrostatic pressure, which acts uniformly on submerged surfaces and is calculated as P = \rho g h, where \rho is the , g is , and h is the depth. Hydrodynamic forces arise from flowing , such as during floods or operations, adding dynamic components to the static hydrostatic load. Seepage forces result from water percolating through the or structure, potentially causing uplift or . Seismic loads introduce inertial forces during earthquakes, requiring consideration of ground acceleration and structure response to prevent . Stability against these loads is evaluated using factors of safety (FS) that compare resisting forces or moments to driving ones, following guidelines such as those from the U.S. Army Corps of Engineers (USACE). For overturning, FS is defined as the ratio of resisting moment to overturning moment and must be at least 2.0 under usual loading conditions to prevent rotation. Sliding stability requires FS of at least 1.5 under usual loading for normal structures, calculated as the ratio of frictional resistance to shear force along the base. Uplift stability addresses buoyant and seepage pressures reducing effective weight, with FS typically set at 1.3 or higher for usual loading, up to 4.0 for critical structures under normal conditions, and decreasing to 1.1 for extreme scenarios depending on the structure type. These criteria ensure the structure remains in equilibrium, with minimum required FS typically higher for normal loading conditions and lower for extreme events to account for the lower probability of such loads while maintaining safety. Material selection is critical for withstanding these loads while minimizing permeability and degradation. , used in and arch dams, offers high (typically 20-40 ) and low permeability when properly mixed with admixtures, enabling it to resist hydrostatic and seismic forces effectively. Earthfill materials, such as compacted clay or rockfill for embankment dams, provide stability through mass and but require low-permeability cores to control seepage. is selected for gates and components due to its high tensile strength and , though coatings are essential to maintain against hydrodynamic . Foundation design focuses on adequate and seepage control to support the structure's weight and loads. Grouting involves injecting or chemical agents into the foundation rock or to fissures, reducing seepage forces and uplift pressures. Ultimate is assessed using Terzaghi's : q_{ult} = c N_c + \gamma D N_q + 0.5 \gamma B N_\gamma where c is , \gamma is unit weight, D is foundation depth, B is width, and N_c, N_q, N_\gamma are factors dependent on angle. This ensures the foundation can resist compressive and stresses without excessive . Ongoing verifies design assumptions and detects potential . Piezometers measure pore water pressures in the and to assess seepage and uplift risks, with data used to adjust operational limits if pressures exceed thresholds. Inclinometers track lateral and vertical settlements by tube deformations, providing early warnings of movement or . These instruments are installed during construction and integrated into automated systems for real-time data collection.

Major Types

Dams and Reservoirs

Dams serve as primary impounding structures in hydraulic engineering, designed to store water in reservoirs for various purposes such as irrigation, hydropower, and water supply. The main types include gravity dams, which rely on the weight of concrete or masonry to resist water pressure; arch dams, which transfer loads primarily through arch action to the abutments; buttress dams, featuring a sloped upstream face supported by triangular buttresses; and embankment dams, constructed from compacted earth or rockfill materials. Selection of the dam type is influenced by site topography, foundation conditions, and available materials; for instance, narrow valleys with strong abutments favor arch dams, while wide valleys with pervious foundations suit embankment types. Key components of dams include the crest, which forms the top elevation of the structure spanning from abutment to abutment and determines the maximum level; the , often featuring an profile to match the lower of a free-falling for efficient overflow discharge; outlet works, consisting of gated conduits typically at the dam to regulate controlled release; and the , a low point or depression in the rim where auxiliary spillways are placed to manage overflow away from the main . These elements ensure safe operation by preventing overtopping and facilitating management. Reservoir management involves strategies to maintain storage capacity, including control through periodic drawdowns that lower water levels to flush accumulated via , thereby preserving usable volume. Drawdown levels are planned elevations for operational purposes, such as seasonal lowering to enhance sediment scour or prepare for . Capacity curves, plotting storage volume against water surface elevation, guide these operations by quantifying active, inactive, and storage zones. Construction of dams proceeds in distinct phases, beginning with foundation preparation that involves excavation, , and of the or to ensure and prevent seepage. For dams, subsequent phases include placement of the impervious core material in thin lifts for compaction, followed by zoning with pervious shells to enhance . methods, used in or dams, involve mixing low-cement and compacting it in layers with vibratory rollers for rapid, economical construction. In arch dams, distribution is analyzed to ensure loads are effectively transferred horizontally to the abutments, with the structure's thin, curved profile relying on compressive arch action rather than mass for resistance. dams employ zoning techniques, such as an impervious central core flanked by semi-pervious shoulders and drainage blankets, to control seepage and prevent by providing filtered paths for water exit while maintaining structural integrity.

Canals and Aqueducts

Canals and aqueducts serve as essential hydraulic structures for the long-distance conveyance of , typically for , , or urban supply, by guiding flow through open channels or elevated conduits while minimizing losses and ensuring structural integrity. These systems differ from impounding structures by emphasizing linear over , often spanning varied terrains with alignments that balance hydraulic efficiency and construction costs. Canals are classified by lining and to suit conditions and . Lined canals feature rigid surfaces such as or geomembranes to reduce seepage losses, which can exceed 50% in unlined channels, thereby improving use efficiency in arid regions. Unlined canals, constructed in or compacted , are more economical for low-permeability terrains but require wider sections to compensate for infiltration. Regarding alignments, canals follow land contours to avoid cross-drainage works, maintaining a near-level grade for uniform flow. canals run along ridge lines or watersheds, dividing drainage basins and necessitating lifts or drops at divides. Side-slope canals traverse hillsides at angles to contours, often requiring frequent cross-drainage to handle perpendicular streams. Aqueducts extend canal principles to elevated or pressurized systems, incorporating features like inverted siphons, flumes, and siphon spillways to navigate obstacles. Inverted siphons are closed conduits that dip below ground or watercourses, operating under to convey across depressions or under drains. Flumes are narrow, elevated open channels, often supported on trestles, for short spans over valleys. spillways provide overflow relief during surges, routing excess water back to the source via siphonic action. Materials commonly include for durability in open sections and steel for pressurized pipes, with corrosion-resistant coatings in saline environments. Design of canals and aqueducts prioritizes hydraulic efficiency through optimized cross-sections, flow equations, and bend configurations. Trapezoidal cross-sections are preferred for their balance of conveyance capacity and excavation volume, achieving near-optimal hydraulic radius with side slopes of 1:1 to 2:1 for lined canals and wider for unlined ones. Flow velocity is calculated using : V = \frac{1}{n} R^{2/3} S^{1/2} where V is average velocity, n is the roughness coefficient (e.g., 0.014 for lining), R is the hydraulic , and S is the channel slope; this empirical relation ensures non-silting, non-scouring flows typically between 1 and 3 m/s. In bends, superelevation—the transverse rise in water surface on the outer bank due to —must be accounted for to prevent overtopping, with design formulas adjusting freeboard by \Delta h = \frac{V^2 b}{g R} where b is the channel top width, R is the to the centerline, and g is . Cross-drainage solutions address intersections with natural streams, with aqueducts carrying the over rivers in open flumes under for free-flow conditions, supported by piers and allowing unobstructed drainage below. In contrast, syphons route the under the river as a closed conduit under , using inverted siphon barrels to maintain full flow without . The , completed in 1869, exemplifies early large-scale conveyance with sections including a 72-foot bottom width and 26-foot depth, incorporating weirs like those at for level regulation across lakes and cuts. Modern examples include the , a 444-mile system with trapezoidal sections (85-foot base, 1.5:1 slopes) and pumping plants like Dos Amigos, which lift over 1,900 feet total via multiple stages to deliver up to 13,100 cfs southward.

Applications and Impacts

Water Resource Management

Hydraulic structures are integral to resource management, facilitating the , diversion, and controlled release of to meet societal needs such as , , and transportation. These structures enable efficient allocation of limited supplies, balancing demands across competing uses while minimizing waste. In systems, diversion weirs serve as low-height barriers across rivers to elevate levels, enabling the extraction of flows into upstream networks for agricultural distribution. These weirs operate by creating a head difference that directs into intake gates connected to main canals, secondary branches, and tertiary distributaries, ultimately delivering it to field application points. networks form the backbone of conveyance, transporting over distances that can span hundreds of kilometers in large schemes, with structures like cross-regulators maintaining equitable flow division. Efficiency in these systems is often quantified by conveyance loss, which encompasses seepage, , and operational spills, typically ranging from 5% to 20% in unlined or semi-lined canals depending on , length, and maintenance practices; lining with or geomembranes can reduce these losses to under 5% in well-managed networks. Flood mitigation relies on hydraulic structures to attenuate extreme flows and protect vulnerable areas. Reservoirs, formed by , perform of inflow hydrographs—graphs depicting incoming flood volumes over time—through storage, producing outflow hydrographs with reduced peaks and extended durations. This process uses the reservoir's elevation-storage-discharge relationship to temporarily impound excess water, releasing it gradually via spillways or outlets to prevent downstream overflow. Complementing reservoirs, systems consist of earthen or concrete embankments aligned parallel to channels, designed to contain floodwaters within defined boundaries and safeguard adjacent floodplains. These s, often reinforced with revetments, can handle floods up to certain magnitudes, such as 100-year events, by raising without altering natural flow paths. Hydropower generation integrates directly into infrastructure to convert energy into . In conventional setups, water from the flows through penstocks—pressurized conduits—to turbine housings embedded in the dam's powerhouse, where rotating blades drive generators. This integration allows multi-purpose to produce while fulfilling storage functions, with turbine types like or Kaplan selected based on head and flow variability. The theoretical output is given by the equation P = \rho g Q H \eta where P is the power (in watts), \rho is the density of water (typically 1000 kg/m³), g is gravitational acceleration (9.81 m/s²), Q is the volumetric flow rate (m³/s), H is the effective head (m), and \eta is the overall efficiency (often 80-90% for modern systems). This formula underscores how higher heads and flows amplify generation capacity, making dams with substantial storage particularly valuable for baseload and peak power. Navigation and municipal further demonstrate the versatility of hydraulic structures in . Locks in systems function as watertight chambers with gated ends, enabling vessels to ascend or descend differences by controlled filling or emptying via culverts and valves, thus maintaining uninterrupted inland on rivers with variable . For , treatment intakes—submerged structures like screened cribs or towers in rivers or reservoirs—regulate the abstraction of , incorporating trash racks and velocity caps to minimize ingress while directing flows to pumping stations for purification. These intakes ensure consistent volumes for urban distribution, often designed to handle seasonal fluctuations in source availability. Basin-wide integrated planning coordinates multiple hydraulic structures to optimize regional water use, as exemplified by the (TVA), established by congressional act in 1933. The TVA oversees a network of 29 dams across the 652-mile system, implementing unified operations for simultaneous flood storage, navigation enhancement through maintained channel depths, , and irrigation support, thereby fostering in a multi-state . This approach demonstrates how centralized management can achieve synergies, such as using releases for downstream while reserving capacity for dry-season .

Environmental and Social Effects

Hydraulic structures, particularly and reservoirs, profoundly alter ecosystems by fragmenting s and disrupting natural river connectivity. physically block riverine pathways, isolating upstream and downstream populations of aquatic species and reducing available habitat size, which intensifies for resources such as spawning grounds and . This fragmentation affects resident and migratory , leading to decreased within populations. Additionally, trap sediments behind reservoirs, preventing their downstream transport and causing in river deltas and coastal areas, which destabilizes habitats and increases vulnerability to and . is severely impeded by these barriers, as structures alter water depths, currents, and temperatures, obstructing access to spawning and feeding areas for species like , eels, and shads, often resulting in population declines. Water quality in reservoirs is significantly altered by hydraulic structures, primarily through processes like and thermal stratification. Reservoirs accumulate nutrients such as from upstream sources, and under anoxic conditions in deeper layers, these nutrients are remobilized, promoting excessive algal blooms that reduce oxygen levels and harm aquatic life. For instance, in tropical reservoirs like , post-flood phosphorus releases have led to invasive weed coverage of 10-15% of the surface area. Thermal stratification, common in most large low-latitude reservoirs, creates distinct layers with cooler, oxygen-depleted hypolimnion water, which, when released, disrupts downstream thermal regimes and delays fish spawning. This stratification exacerbates and nutrient cycling imbalances, further degrading water quality and ecosystem health. Social consequences of hydraulic structures often include large-scale community displacement and loss. The in , for example, displaced over 1.3 million people from 1,711 villages, 356 communes, 116 towns, and 20 cities, leading to economic uncertainty, social conflicts, and separation among resettled families. Resettlement has exacerbated , particularly among women and farmers who lost 34,000 hectares of , while host communities have ostracized newcomers, increasing risks of diseases like . Culturally, inundation has submerged archaeological sites and sacred locations, eroding and historical identities in affected regions. Sustainability challenges for hydraulic structures are amplified by , which necessitates adaptive designs to counter heightened risks. Projected increases in rainfall intensity and altered storm patterns can elevate magnitudes by up to 35% due to temporal shifts alone, overwhelming dam storage and capacities in urban watersheds. Combined with volume increases, such as 208 mm over 24 hours versus baseline 160 mm, risks may rise by 10% to 170%, requiring resilient infrastructure modifications. Decommissioning offers a pathway to restoration; the dams' removal from 2011 to 2014 released 20 million tons of , enhancing nearshore habitats for like sand lance and geoducks while allowing rapid recovery in cleared waters. Mitigation strategies emphasize proactive environmental impact assessments (EIAs) and to minimize these effects. EIAs for projects integrate environmental and social impact assessments, using decision trees to select assessment resolutions based on sensitivity and social reliance, ensuring sustainable flow regimes through hydrological and ecological . involves ongoing monitoring of ecosystem responses, , and operational adjustments via environmental flow management plans, allowing refinements to releases if targets for or are unmet. Fish ladders, as partial mitigations, facilitate upstream passage for migratory , though their effectiveness varies by and conditions.

Maintenance and Safety

Inspection Methods

Inspection methods for hydraulic structures encompass a range of techniques designed to detect structural distress, monitor performance, and ensure long-term safety without compromising the integrity of the facilities. These methods are essential for identifying issues such as cracks, seepage, and material degradation in , canals, aqueducts, and associated components like and spillways. Visual and geophysical surveys form the foundation of routine assessments, providing initial insights into surface and subsurface conditions. Visual surveys involve systematic on-site examinations to map cracks, erosion, and seepage patterns on exposed surfaces of , , or elements. For instance, mapping documents the , , and of fissures using photographs and sketches, which can indicate settlement or overstressing in dams and canal linings. Geophysical surveys complement these by employing non-invasive tools to probe subsurface anomalies; (GPR) is particularly effective for detecting voids, features, or in embankments and aqueduct foundations, offering resolution down to depths of several meters. , a key geophysical technique, assesses integrity by measuring wave propagation speeds to identify internal cracks or voids, often applied to structures and canal walls. Instrumentation enables continuous or periodic monitoring of dynamic behaviors in hydraulic structures. Strain gauges measure deformation in concrete dams and steel gates, providing data on stress distribution under varying loads. Tiltmeters detect angular changes in structure alignment, crucial for identifying in canal banks or reservoir embankments. Supervisory Control and Data Acquisition () systems integrate these sensors for real-time data collection and alerting, facilitating remote oversight of multiple sites like aqueduct networks. Piezometers and observation wells track pore water pressures to monitor internal stability. Hydraulic performance checks verify the of water conveyance and control features. Flow measurements using acoustic Doppler profilers (ADCPs) or meters quantify in , spillways, and outlets, ensuring capacities align with design specifications during peak flows. Seepage monitoring employs weirs and flumes to measure leakage rates from reservoirs or linings, with elevated or indicating potential risks; further delineates leak paths by injecting fluorescent tracers and observing downstream emergence. These checks are vital for preventing in unlined channels or gated structures. Non-destructive testing (NDT) methods extend assessments to material properties without invasive procedures. Beyond , techniques like for gates in locks and dye penetrant testing reveal surface-breaking flaws in welds and coatings on aqueduct supports. and resistivity surveys detect subsurface moisture anomalies indicative of leaks in embankment dams or canal beds. These methods adhere to standards such as ASTM E709 for magnetic particle testing, ensuring reliable detection of defects as small as 0.1 mm. Inspection frequencies and standards are governed by risk-based protocols to prioritize high-hazard structures. The International Commission on Large Dams (ICOLD) recommends annual comprehensive inspections for large dams, with more frequent informal checks during operations. U.S. Army Corps of Engineers guidelines specify periodic inspections every 1-5 years for dams, supplemented by post-flood evaluations to assess scour or displacement. For canals and aqueducts, routine visual checks occur daily or monthly, with detailed NDT every 2-3 years or after seismic events, aligning with federal standards like those in ER 1110-2-1156. These protocols ensure proactive maintenance, reducing failure risks through standardized reporting and multidisciplinary team involvement.

Failure Case Studies

The , an earthfill structure in southeastern , , failed catastrophically on June 5, 1976, during its first filling, releasing approximately 310,000 acre-feet of water and causing widespread flooding along the Teton River. The primary cause was internal erosion, or , initiated by seepage through fractures in the underlying foundation and right , exacerbated by inadequate grouting and walls that failed to seal the permeable zones adequately. This seepage led to progressive erosion of the material, culminating in a about 300 feet wide; the disaster resulted in 11 deaths and approximately $400 million to $1 billion (1976 dollars) in damages to property, , and . In August 1975, the in Province, , collapsed following extreme rainfall from Typhoon Nina, which dumped up to 1,062 mm of rain in a single day, far exceeding the dam's design capacity of 300 mm. The failure occurred via overtopping when the reached 116.34 meters, triggering a of breaches in 62 downstream dams and reservoirs, amplifying the floodwave to affect over 11 million people across 29 counties. flaws, including insufficient capacity and inadequate maintenance during the storm, combined with communication breakdowns that delayed evacuation warnings, led to more than 150,000 deaths from , subsequent , and disease in the ensuing weeks. The disaster on October 9, 1963, in , demonstrated the perils of geohazards despite a structurally sound thin-arch that withstood the event intact. A massive of about 270 million cubic meters of rock detached from Mount Toc and plunged into the at speeds up to 30 meters per second, displacing water to generate an overflow wave up to 250 meters high that overtopped the 261-meter by 70 meters. This surge devastated villages in the Piave Valley below, killing approximately 2,000 people, with the death toll concentrated in where nearly the entire population of 1,459 perished; prior warnings of instability were downplayed, and drawdown was insufficient to mitigate the wave's . In February 2017, the in , , experienced a major failure during heavy rainfall, leading to emergency evacuations of nearly 200,000 people downstream. The incident was caused by and damage to the concrete , exacerbated by inadequate maintenance and design flaws that allowed water to undermine the structure; no deaths occurred, but repairs cost over $1 billion (2017 dollars) and highlighted vulnerabilities in aging infrastructure to . Post-event investigations emphasized improved and regular inspections to prevent similar near-misses. These failures underscore the critical need for (PRA) in hydraulic structure design and operation, which quantifies uncertainties in loading, material behavior, and failure modes to prioritize mitigation over deterministic approaches alone. PRA, informed by post-event analyses like those of Teton and Vajont, evaluates the likelihood and consequences of rare events such as extreme storms or landslides, enabling better-informed decisions on sizing and treatment. Additionally, the of emergency action plans (EAPs) has become standard, outlining detection, notification, and response protocols to evacuate downstream populations within hours of an impending breach, as evidenced by simulations from Banqiao's communication failures. In response to these incidents, modern preventions include enhanced regulatory frameworks, such as the U.S. Army Corps of Engineers' (USACE) Dam Safety Program, established and refined post-1976 to mandate periodic risk-informed assessments and remediation for all federal dams. USACE guidelines now require potential failure mode analyses (PFMAs) to identify and address vulnerabilities like or overtopping, integrated with inundation mapping and EAP exercises, reducing overall failure risks through a portfolio approach that considers aging and variability.

References

  1. [1]
    Hydraulic Structure - an overview | ScienceDirect Topics
    Hydraulic structures are engineered constructions designed to manage and control water flow, including retaining, conveying, or disrupting natural flow.Missing: civil | Show results with:civil
  2. [2]
    [PDF] Design of Hydraulic Structures 89 - Bureau of Reclamation
    Hydraulic models are used to develop and verify designs of waterways, including spillways, and to fine-tune details for cost savings.
  3. [3]
    None
    ### Summary of Hydraulic Structures Content
  4. [4]
  5. [5]
    None
    Summary of each segment:
  6. [6]
    Section 2: Introduction to Hydraulic Analysis and Design
    The hydraulic design or analysis of highway drainage facilities usually involves a general procedure, the specific components of which vary for each project.Missing: civil | Show results with:civil
  7. [7]
    Hydraulic Head - Lower Colorado Region - Bureau of Reclamation
    Oct 15, 2021 · “Hydraulic head,” or simply “head,” is the mechanical energy per unit weight of groundwater. Groundwater flows from positions of high hydraulic ...
  8. [8]
    Discharge - Texas Department of Transportation
    In its simplest concept discharge means outflow; therefore, the use of this term is not restricted as to course or location, and it can be applied to ...
  9. [9]
    [PDF] WEIR EXPERIMENTS, COEFFICIENTS, AND FORMULAS
    INTRODUCTION. DEFINITIONS OF TERMS. The word "weir" will be used to describe any structure used to. determine the volume of flow of water from measurements of ...
  10. [10]
    [PDF] Early Hydraulic Civilization in Egypt
    This book studies the cultural ecology of early hydraulic civilization in Egypt, focusing on the rise of an "irrigation civilization" and its factors.
  11. [11]
    The Aqueducts and Water Supply of Ancient Rome - PMC
    The first aqueduct was the Aqua Appia, erected in 312 BC by the censor Appius Claudius Caecus (c. 340 to 273 BC). During the Republican period, three more ...
  12. [12]
    Water Management and Hydraulic Technology - Muslim Heritage
    Dec 30, 2001 · Muslims developed hydraulic technology, including dams, reservoirs, and water management for agriculture, and used dams for water storage and ...
  13. [13]
    [PDF] SANCTIFIED SCIENCE - Medieval monks developed technologies ...
    Jun 1, 2025 · The. Cistercians employed water in metal- lurgy, using waterpower to lift ore out of mines, pulverize it in hammer mills, and then use water- ...
  14. [14]
    Navigli: Leonardo da Vinci in Milan: guided tours
    In 1100s Milan had already developed a circular moat system used for irrigation purposes and to operate hydraulic wheels to mill the wheat, cut woods etc.Missing: medieval Islamic European
  15. [15]
    Historical Development of Arch Dams. From Cut-Stone Arches to ...
    Jan 28, 2023 · Arch dam development includes Roman, Mongol, early 19th century, Australian concrete, and modern designs. The first was likely the Roman dam at ...
  16. [16]
    The First Aswan Dam
    In 1899, construction of the first Aswan Dam was begun. Completed in 1902, its height was raised in subsequent building campaigns of 1907-12 and 1929-34.
  17. [17]
    John Smeaton - ASCE
    John Smeaton was born in Austhorpe Lodge near Leeds, England, on June 8, 1724. John Smeaton's prolific career mirrored the developments of the Industrial ...Missing: weir | Show results with:weir
  18. [18]
    Hoover Dam | Bureau of Reclamation
    Mar 12, 2015 · When Hoover Dam was finished in 1935 it was the tallest dam in the world. From about 1938 until 1948 the Hoover Dam powerplant was the ...Missing: 1936 2003
  19. [19]
    [PDF] Geographical Overview of the Three Gorges Dam and Reservoir ...
    after the reservoir began filling in 2003 (There had had been a M = 6 in the past). Three Gorges Dam is built to withstand a M = 7.0 earthquake. Page 59 ...
  20. [20]
    [PDF] Highlights in the History of Hydraulics
    Even Franklin was not the first to conduct scale-model tests, credit for which is due John Smeaton (1724-92), an English engineer who was one of the very few ...
  21. [21]
  22. [22]
    [PDF] HYDRAULIC STRUCTURES
    Based on use, dams are classified as follows i) Storage dam,. (ii) Diversion dam,. (iii) Detention dam. Storage dam: • This is the most common type of dam ...
  23. [23]
    [PDF] Hydraulic Design of Stilling Basins and Energy Dissipators
    THIS MONOGRAPH generalizes the design of stilling basins, energy dissipators of several kinds and associated appurtenances. General design rules are presented ...
  24. [24]
    [PDF] Small hydraulic structures - FAO Knowledge Repository
    The width (or breadth) of the weir crest (Bt) should not be less than. 0. 30 ... measurement is less if the height of the weir crest above the bottom of the ...
  25. [25]
    [PDF] Discharge measurement structures - SamSamWater Foundation
    It is intended to serve as a guide to good practice for engineers concerned with the design and operation of such structures. It is hoped that the book will ...
  26. [26]
    ICOLD CIGB > General Synthesis
    Definition of a “Large Dam”. A dam with a height of 15 metres or greater from lowest foundation to crest or a dam between 5 metres and 15 metres impounding ...
  27. [27]
    ICOLD CIGB > Definition of a Large Dam
    A dam with a height of 15 metres or greater from lowest foundation to crest or a dam between 5 metres and 15 metres impounding more than 3 million cubic metres.Missing: small | Show results with:small
  28. [28]
    Typical layout of a barrage in a diversion head works - ResearchGate
    Barrages are hydraulic structures constructed across rivers to divert flow into irrigation canals or power generation channels. The most of these structures are ...
  29. [29]
    Seawalls and revetments - Coastal Wiki
    Jan 19, 2025 · Seawalls or revetments are shore-parallel structures at the transition between the (sandy) beach and the (higher) mainland or dune.Introduction · Solving coastal engineering... · Scour at the toe of the seawall...
  30. [30]
    Effectiveness Of Flood Barriers In Urban Environments
    How flood barriers protect urban areas, explore their types and effectiveness, and learn how strategic placement and community involvement enhance flood ...
  31. [31]
    What Factors Influence Geological Site Selection? → Question
    Apr 23, 2025 · Geological site selection depends on subsurface structure, water presence, stability, and integrating human dimensions for sustainable ...Missing: volume | Show results with:volume
  32. [32]
    Development of Site-Scale Conceptual Model Using Integrated ...
    Apr 20, 2022 · Additionally, geological structures such as fractures, faults, and deformation bands are also considered important factors affecting fluid flow ...3. Results · 3.1. Lithological... · 3.4. Groundwater Flow...
  33. [33]
    [PDF] Tier 2 - Chapter 03 - Design of Struct., Comp., Equip. and Systems - S
    The the lateral hydrostatic and hydrodynamic pressure on the structures due to the design flood water level, as well as ground and soil pressures, are ...Missing: hydraulic | Show results with:hydraulic
  34. [34]
    [PDF] CHAPTER III GRAVITY DAMS
    The seepage potential is calculated by first determining the creep distance which a molecule of water would follow as it flows beneath the structure.
  35. [35]
    [PDF] Earthquake Design and Evaluation of Concrete Hydraulic Structures
    May 1, 2007 · Purpose. This manual provides guidance for performance-based design and evaluation of concrete hydraulic structures (CHS).
  36. [36]
    [PDF] stability analysis of concrete structures - USACE Publications
    Dec 1, 2005 · The types of hydraulic structures, which could be classified as critical, include gravity dams and spillways, arch dams, urban flood walls ...
  37. [37]
    [PDF] EM 1110-2-2000 - USACE Publications
    Mar 31, 2001 · Emphasis is placed on the problems of concrete for hydraulic structures. Roller-compacted concrete, shotcrete, rigid pavement, architectural ...
  38. [38]
    [PDF] Materials for Embankment Dams
    Embankment dam materials include soil, rockfill, granular filters, asphalt concrete, concrete facing, geosynthetics, reinforced fill, slope protection, and ...
  39. [39]
  40. [40]
    Using Terzaghi's Equation in Foundation Design - Geoengineer.org
    Nov 25, 2022 · The design approach used for helical piles and anchors makes use of the basic General Bearing Capacity Equation (GBCE), which was developed in the 1940s by ...
  41. [41]
    Introduction to Foundation Engineering using Terzaghi's Bearing ...
    Oct 27, 2023 · Terzaghi's general bearing capacity equation for determining ultimate bearing capacity, as given in most Foundation Engineering textbooks, may be stated as:
  42. [42]
    [PDF] CHAPTER IX INSTRUMENTATION AND MONITORING
    Chapter IX covers instrumentation and monitoring, including an introduction, philosophy, and types of instrumentation such as water level and pressure.
  43. [43]
    [PDF] Embankment Dams - Bureau of Reclamation
    Oct 4, 2011 · In this design standard, three broad classifications are used: purpose, hydraulic design, and materials comprising the structure.
  44. [44]
    [PDF] General Design and Construction Considerations for Earth and ...
    Jul 30, 2004 · a. Introduction. The two principal types of embankment dams are earth and rock-fill dams, depending on the predominant fill material used.
  45. [45]
    [PDF] Embankment Dams - Bureau of Reclamation
    An earthfill dam is designed considering the topographic and foundation conditions at the site and using available construction materials. 2.2.2.1.1 Diaphragm ...
  46. [46]
    [PDF] 2021-NWP-53: Removal of Low-Head Dams
    Feb 25, 2022 · The dam crest is the top of the dam from left a butment to right a butment. A low- hea d dam may have been built for a ra nge of purposes (e.g.,.
  47. [47]
    [PDF] Hydraulic Design of Spillways - USACE Publications
    Aug 31, 1992 · The ogee crest spillway is basically a sharp-crested weir with the ... and Outlet Works, Townshend Dam and Reservoir, West River, Vermont;.Missing: components | Show results with:components
  48. [48]
    New England District > Missions > Civil Works > Glossary
    Outlet Works gated conduits, usually located at the base of a dam, that regulate the discharge of water. Pumping Station a structure containing pumps that ...
  49. [49]
    Reclamation Projects & Facilities Glossary
    Apr 25, 2017 · Arch-buttress dam (or curved buttress dam). A buttress dam which is curved in plan. Arch-gravity dam. An arch dam which is only slightly thinner ...
  50. [50]
    [PDF] Reservoir Evolution, Downstream Sediment Transport
    Nov 7, 2018 · storage curves and suggests rates of reservoir sedimentation have been low. This is consistent with studies scoping the viability and ...<|control11|><|separator|>
  51. [51]
    [PDF] Best Practices of Numerically Modeling Hydraulic Flushing of ...
    Sep 30, 2022 · The document provides best practices in numerically simulating hydraulic flushing for reservoir sediment management. Three types of hydraulic ...
  52. [52]
    Reservoir evolution, downstream sediment transport, downstream ...
    May 17, 2024 · Drawdown: This term generally refers to the dam management action of lowering pool levels to different elevations to meet a specific purpose.
  53. [53]
    [PDF] Embankment Dams - Bureau of Reclamation
    May 16, 2012 · Preparation of the foundation and abutments for an earth or rockfill dam is a critical phase of construction; the thoroughness with which it is ...
  54. [54]
    [PDF] Foundation Preparation, Treatment, and Cleanup
    This chapter discusses foundation preparation for and the placement of the first several layers of earthfill and concrete. Preparation includes excavating ...Missing: phases | Show results with:phases
  55. [55]
    [PDF] EM 1110-2-2006 - USACE Publications
    Jan 15, 2000 · Purpose. The purpose of this manual is to provide information and guidance on the use of roller- compacted concrete (RCC) in dams and other ...
  56. [56]
    [PDF] EM 1110-2-2201 - USACE Publications
    May 31, 1994 · This manual provides information and guidance on the design, analysis, and construction of concrete arch dams. 2. Applicability. This manual ...<|control11|><|separator|>
  57. [57]
    [PDF] seepage analysis and control for dams - USACE Publications
    Apr 30, 1993 · Seepage control is necessary to prevent excessive uplift pressures, sloughing of the downstream slope, piping through the embankment and ...
  58. [58]
    [PDF] Design Standards No. 3, "Canals and Related Structures"
    Design Standards No. 3 covers canals, laterals, diversion dams, headworks, water measurement, cross drainage, pipe systems, and bridges. It emphasizes lining ...
  59. [59]
    [PDF] Classification of Canals based on Different Factors
    A canal may be aligned as a contour canal, a side slope canal or a ridge canal according to the type of terrain and culturable area. Irrigation canals can ...Missing: summit | Show results with:summit
  60. [60]
    [PDF] Irrigation canals can be aligned in any of the three ways
    Irrigation canals can be aligned as watershed, contour, or side slope canals. Side slope canals are aligned at right angles to contours.Missing: summit | Show results with:summit
  61. [61]
    (PDF) Most Hydraulically Efficient Standard Lined Canal Sections
    Aug 7, 2025 · Cross-sectional dimensions of the most hydraulically efficient lined canals are evaluated based on an analysis of a generalized trapezoidal ...
  62. [62]
    Manning's Equation
    Q = Flow Rate, (ft3/s). v = Velocity, (ft/s). A = Flow Area, (ft2). n = Manning's Roughness Coefficient. R = Hydraulic Radius, (ft). S = Channel Slope, (ft/ft).
  63. [63]
    [PDF] Design of Roadside Channels with Flexible Linings
    A final consideration for channel design at bends is the increase in water surface elevation at the outside of the bend caused by the superelevation of the ...
  64. [64]
    [PDF] cross drainage works | JITS
    Syphon Aqueduct:​​ drainage, but the drainage water cannot pass clearly below the canal. It flows under siphonic action. So, it is known as siphon aqueduct.Missing: rivers | Show results with:rivers
  65. [65]
    [PDF] The construction of the Suez Canal - WIT Press
    3.4 Hydraulic engineering​​ The filling of the canal and its lakes was regulated through El Timsah, Serapeum, and El Suez Weirs. The final filling of the Canal ...
  66. [66]
    [PDF] California Aqueduct Hydraulic Model Development
    This report is about the California Aqueduct Hydraulic Model Development, prepared by the California Department of Water Resources in December 2023.
  67. [67]
    Diversion Weirs - International Commission on Irrigation and Drainage
    A diversion weir is a low obstruction that raises water levels to divert water when flow rates exceed a threshold. The diversion rate depends on the water ...
  68. [68]
    CHAPTER 5 - IRRIGATION SYSTEM
    The conveyance system assures the transport of water from the main intake structure or main pumping station up to the field ditches. The distribution system ...
  69. [69]
    A National Estimate of Irrigation Canal Lining and Piping Water ...
    The average irrigation delivery organization reports a conveyance loss of 15 percent of the total water brought into their system in 2019. Using a control ...
  70. [70]
    [PDF] CHAPTER 10, FLOW ROUTING, HANDBOOK OF HYDROLOGY, (ED
    Mar 10, 2023 · Flow routing is a mathematical procedure for predicting the changing magnitude, speed, and shape of a flood wave as a function of time.Missing: mitigation | Show results with:mitigation
  71. [71]
    [PDF] Chapter 17 Flood Routing
    Definitions related to reservoirs and breaching dams are: Routing, reservoir—This procedure derives the out- flow hydrograph from a reservoir from the inflow.Missing: mitigation | Show results with:mitigation
  72. [72]
    Living With Levees | FEMA.gov
    May 13, 2024 · The United States has thousands of miles of levee systems built to help contain or control the flow of water to reduce the risk of flooding.
  73. [73]
    Hydropower Does More Than You Think: Six Things To Know About ...
    Sep 15, 2022 · Hydropower plants produce energy using the elevation difference created by a dam or diversion structure. Water flows in one side and exits ...
  74. [74]
    How a Lock Works - Chesapeake & Ohio Canal National Historical ...
    Dec 10, 2022 · A lock is an elevator for boats. It lifts boats up or down to the next level on the canal. There are 74 locks on the C and O Canal.
  75. [75]
    7. Main Water Intake Structures
    Main water intakes are used for the overall regulation and diversion of water supplies to a pond or group of ponds.
  76. [76]
    Our History - Tennessee Valley Authority
    Signed in 1933, the Tennessee Valley Authority Act created a public corporation “To improve the navigability and to provide for the flood control of the ...Missing: basin- wide
  77. [77]
    [PDF] Implications of Dam Obstruction for Global Freshwater Fish Diversity
    2007). Dams inhibit organism migration, alter and fragment habitats, and reduce resource transport throughout rivers. (Roberts 2001, Hall et al. 2011), thus ...
  78. [78]
    Dams, water quality and tropical reservoir stratification - BG
    Apr 23, 2019 · In this paper, we review and synthesize information on the effects of damming on water quality with a special focus on low latitudes.
  79. [79]
    [PDF] China's Three Gorges Dam: Development, Displacement, and ...
    May 26, 2021 · This writing serves to discuss the environmental and social impacts that have occurred throughout the development of the Three Gorges Project, ...
  80. [80]
    Increase in flood risk resulting from climate change in a developed ...
    Mar 29, 2018 · Climate change causes more frequent extreme rainfall, increasing flood risk. Temporal pattern changes increase flood risk up to 35%, and ...
  81. [81]
    Elwha River: New Study Examines Effects of Dam Removals on ...
    Sep 7, 2023 · Dam removal has gained traction as a powerful tool for restoring aquatic habitats and eliminating high-risk infrastructure. While most previous ...
  82. [82]
    [PDF] Environmental Flows for Hydropower Projects
    This Good Practice Handbook provides guidance to practitioners on selecting an appropriate environmental flows assessment level for hydropower project.
  83. [83]
  84. [84]
    [PDF] Safety Evaluation of Existing Dams - Bureau of Reclamation
    -Concrete dams encompass a variety of structures which include gravity, slab and buttress, multiple arch, and single arch dams. Regardless of the type of dam, ...
  85. [85]
    [PDF] Inspection, Evaluation, and Repair of Hydraulic Steel Structures
    Dec 1, 2001 · This manual describes the inspection, evaluation, and repair of hydraulic steel structures, including preinspection identification of critical ...
  86. [86]
    Operation and Maintenance of Hydraulic Structures - IntechOpen
    The destructive examinations may include (i) hydraulic pumping tests for porosity and (ii) permeability and leak detection through a physical and chemical test ...
  87. [87]
    ICOLD CIGB > Dams Safety
    The structure must be continually supervised and inspected throughout its whole life, to ensure that it remains in good health. Most Frequent Causes of Dam ...
  88. [88]
    [PDF] Failure of Teton Dam by Independent Panel to Review Cause of ...
    Dec 31, 1976 · Teton Dam failed on June 5, 1976, when the reservoir was at El. 5301.7, 3.3_ ft below the spillway sill. Although downstream warnings are ...
  89. [89]
    Tragedy at Teton: 1976 Dam Break Disaster - NOAA VLab
    Jun 3, 2020 · In total, there were 11 reported deaths, and the flood caused more than $2 billion in damages to property and infrastructure.
  90. [90]
    [PDF] The World's Most Catastrophic - Dam Failures
    In August 1975, a typhoon struck Zhumadian Prefecture of Henan Prov- ince in central China, causing reservoirs to swell with rainwater behind dozens of dams.* ...
  91. [91]
    Typhoon Nina and the August 1975 Flood over Central China in
    The collapse of the Banqiao and Shimantan Dams initiated cascading failures of 62 smaller dams in southern Henan Province. Major floods occurred over the ...<|separator|>
  92. [92]
    Typhoon Nina–Banqiao dam failure - Britannica
    Oct 10, 2025 · The ensuing floods caused more than 150,000 casualties, making it one of the deadliest typhoon disasters in history.
  93. [93]
    [PDF] VAJONT DAM - ASDSO Lessons Learned
    Officials thought the reservoir was drawn down enough to not cause harm, but they grossly underestimated the resultant size of the landslide. Warnings were not ...
  94. [94]
    Expecting Disaster: The 1963 Landslide of the Vajont Dam
    On October 9, 1963, a landslide above the Vajont Dam created a wave that destroyed several villages in the valley, killing about 2000 people.
  95. [95]
    [PDF] Achieving Public Protection with Dam Safety Risk Assessment ...
    Risk assessment uses guidelines, identifies risks, and helps understand uncertainties. It recognizes all dams have risk, identifies significant factors, and ...Missing: lessons probabilistic
  96. [96]
    Emergency Action Planning - Association of State Dam Safety Officials
    Lessons Learned from Dam Failures and Incidents Website (DamFailures.org) ... Emergency Action Plans (EAPs) help dam owners as well as people downstream.
  97. [97]
    ER 1110-2-1156 - USACE Publications
    Mar 31, 2014 · ... Dam Safety Program - Introduction, Overview, and Guiding Principles a. Added a definition for the term 'Dam'. b. Principles for Dam Safety ...<|separator|>