Fact-checked by Grok 2 weeks ago

Open-channel flow

Open-channel flow is the movement of a , such as , with a exposed to the atmosphere in an open conduit, where the flow is primarily driven by rather than . This contrasts with closed-conduit flows, like those in , where the liquid is fully enclosed and pressure dominates; in open channels, the is deformable, experiences zero , and allows to shape the flow geometry. Common examples include rivers, canals, ditches, and systems, where the flow interacts with along the surface. The behavior of open-channel is characterized by its classification into and nonuniform types. Uniform occurs when the depth and remain constant along the length, typically in prismatic channels with constant and roughness, balancing gravitational driving force with frictional resistance. Nonuniform , on the other hand, features varying depth and , subdivided into gradually varied (where changes occur slowly due to mild alterations or obstructions) and rapidly varied (such as hydraulic jumps from abrupt transitions). These classifications are essential for hydraulic design and analysis in applications. Governing principles for open-channel flow derive from fundamental equations, including the for mass , the momentum equation for force balances, and the energy equation for total head , adapted to account for the free surface. For uniform flow, the empirical Manning's equation is widely used to relate average velocity V to channel properties: V = \frac{1}{n} R_h^{2/3} S_0^{1/2}, where n is the Manning roughness coefficient, R_h is the hydraulic radius, and S_0 is the bed slope. The (Fr = \frac{V}{\sqrt{g y}}, with y as flow depth and g as ) further delineates flow regimes: subcritical (Fr < 1, tranquil, wave-like), critical (Fr = 1, transitional), and supercritical (Fr > 1, shooting, rapid). These concepts underpin practical computations for , , and waterway design.

Fundamentals

Definition and characteristics

Open-channel flow refers to the flow of a , typically , in a conduit or where the upper surface of the is exposed to , resulting in a that partially fills the cross-section. This configuration contrasts with fully enclosed flows, as the does not completely occupy the conduit, allowing the surface to adjust freely to changes in flow conditions. A defining characteristic of open-channel flow is its reliance on as the primary driving force, with the maintained at constant , enabling the flow to respond dynamically to slope and . The flow depth and hydraulic radius—defined as the cross-sectional area divided by the wetted perimeter—vary spatially and temporally, influenced by channel , roughness, and rates. These properties make open-channel flow particularly sensitive to external factors like bed slope and obstructions, often leading to complex surface profiles. Fluid properties such as , which governs the gravitational component, and , which affects resistance, play foundational roles, though they are generally less dominant in high-Reynolds-number flows compared to geometric influences. Examples of open-channel flow abound in both natural and engineered systems, including rivers and streams where occurs naturally, irrigation canals designed for controlled water distribution, spillways that manage excess reservoir outflow, and stormwater drainage channels in urban areas. The historical development of open-channel flow analysis traces back to 19th-century efforts, notably by Henri Darcy, who conducted experiments on hydraulic resistance applicable to channels for systems, and Robert Manning, who formulated empirical relations for practical use in Irish public works focused on and flood mitigation.

Differences from pressurized flow

Open-channel flow differs fundamentally from pressurized flow in closed conduits, such as , primarily due to the presence of a exposed to the atmosphere. In pressurized systems, the completely fills the conduit, and flow is driven by a imposed along the length, whereas open-channel flow relies on acting along the channel to propel the , with the water depth adjusting freely to balance forces. This distinction arises because open-channel flow occurs in partially filled channels where the upper boundary is an air-water interface at constant , contrasting with the rigid, enclosed boundaries of that confine the under varying internal pressures. The distribution in open-channel flow is hydrostatic and varies linearly with depth below the , where the equals atmospheric at the surface and increases proportionally with the submergence of any point in the cross-section. In contrast, pressurized exhibits a more uniform across the cross-section perpendicular to the direction, with the primary variation occurring along the conduit due to and imposed gradients, without a reference. This hydrostatic nature in open channels simplifies certain analyses but introduces variability tied to depth, unlike the consistent enforcement in full pipes. Boundary conditions further highlight this divergence: the in open-channel maintains constant atmospheric , eliminating any confining from surrounding walls, while is bounded entirely by solid surfaces that sustain the internal and dictate the path rigidly. Velocity profiles in open-channel flow tend to be more uniform across the depth compared to pressurized , owing to reduced shear influence from the , which allows larger eddies to form and mix more effectively away from the bed. In laminar conditions, both exhibit parabolic profiles, but in turbulent regimes—common in practical flows—open-channel profiles become blunter near the surface with a sharper adjacent to the solid , differing from the more symmetric, wall-dominated profiles in where is constrained by opposing boundaries. These profile differences stem from the distinct structures: open-channel flow features eddies that interact primarily with the bed and , while involves cross-stream eddy interactions that promote greater uniformity toward the center. Practically, these differences make open-channel flow easier to observe and measure, as surface elevations and depths can be directly visualized and gauged without invasive tools, facilitating applications in natural systems like and engineered open conduits such as aqueducts. In pressurized pipelines, monitoring requires pressure sensors and flow meters embedded in the system, suiting closed infrastructure like networks but complicating direct assessment. This contrast influences design choices: open channels leverage gravity for low-maintenance transport in or , while pipes enable pressurized delivery over varied terrains but demand pumps to overcome friction losses.

Flow Classifications

Geometric classifications

Open-channel flows are geometrically classified based on the shape and configuration of the channel cross-section, which influences flow resistance, capacity, and stability. Natural channels, such as meandering rivers and streams, typically feature irregular cross-sections formed by and , often approximating parabolic or trapezoidal shapes with variable bed roughness due to boulders, , and deposits. In contrast, artificial channels, including canals, flumes, and culverts, are engineered with regular cross-sections like rectangular, trapezoidal, or circular forms to optimize conveyance and minimize . Channels are further distinguished as prismatic or non-prismatic depending on the variation in cross-section along the flow direction. Prismatic channels maintain a constant cross-sectional shape and bed slope over their length, simplifying hydraulic computations for uniform flow conditions. Non-prismatic channels, however, exhibit varying , such as widening or narrowing sections in spillways or natural confluences, which introduce complexities in flow distribution and energy dissipation. The distinction between wide and narrow channels hinges on the width-to-depth , affecting velocity profiles and hydraulic approximations. Wide channels, where the width exceeds approximately 10 times the flow depth, allow for assumptions in the central portion, with the hydraulic closely approximating the flow depth. Narrow channels, with lower width-to-depth s, exhibit more pronounced three-dimensional effects and lateral velocity variations. Lined and unlined channels represent another geometric consideration, primarily impacting and potential. Lined channels, constructed with materials like or stone , provide smooth, stable surfaces that reduce losses, whereas unlined channels, often earth or vegetated banks in natural or trapezoidal sections, experience higher roughness from erodibility and growth. Parabolic cross-sections, common in natural streams and grassed waterways, exemplify how can balance conveyance efficiency with . These geometric features collectively shape the behavior by dictating wetted perimeter and flow resistance.

Regime-based classifications

Open-channel flows are classified into regimes based on dynamic characteristics, primarily using dimensionless numbers to characterize the balance between inertial, gravitational, and viscous forces. The serves as the key parameter for distinguishing between subcritical, critical, and supercritical regimes, reflecting the relative importance of flow inertia to . Additionally, the delineates laminar from turbulent flow, though turbulent conditions predominate in most practical open-channel scenarios. The , denoted as Fr, is defined as the ratio of the mean to the speed of a shallow-water : Fr = \frac{V}{\sqrt{g D}} where V is the mean of the (in m/s), g is the (approximately 9.81 m/s²), and D is the hydraulic depth, calculated as the cross-sectional area of the divided by the top width at the surface (D = A / T, in meters). This dimensionless parameter (Fr has no units) quantifies the flow's tendency to support or propagate surface waves, with values determining the regime. Subcritical flow occurs when Fr < 1, characterized by relatively low velocities and greater depths compared to critical conditions, resulting in a tranquil regime where disturbances propagate both upstream and downstream like waves on a pond. In this regime, the flow depth exceeds the critical depth, and gravitational forces dominate over inertial ones, allowing downstream controls (such as weirs) to influence upstream conditions. Supercritical flow arises when Fr > 1, featuring high velocities and shallow depths relative to critical conditions, often described as a shooting or rapid flow where inertial forces prevail, and disturbances cannot propagate upstream against the current. Here, the flow depth is less than the critical depth, making the regime analogous to high-speed shallow water where upstream influences are negligible, and control is exerted from upstream structures. Critical flow corresponds to Fr = 1, representing a transitional state where inertial and gravitational forces are balanced, occurring at the minimum for a given and thus the point of maximum efficiency for that energy level. This regime is unstable and typically forms at locations of , such as immediately downstream of a sluice gate or at the of a broad-crested , where the accelerates to this threshold before transitioning to supercritical conditions. Classification by viscosity effects uses the Reynolds number, Re = \frac{V R}{\nu}, where R is the hydraulic radius (cross-sectional area divided by wetted perimeter, in meters), and \nu is the kinematic viscosity of water (approximately 1.0 × 10^{-6} m²/s at 20°C). Flows with Re < 500 are laminar, featuring smooth, orderly motion dominated by viscous forces; Re > 2000 indicates turbulent flow with chaotic eddies and mixing; and intermediate values represent transitional regimes. However, in open channels, turbulent flow is dominant due to the typically high velocities, rough boundaries, and large scales encountered in natural and engineered systems, rendering laminar conditions rare except in very slow, shallow flows.

Flow States

Uniform flow

Uniform flow represents a steady, one-dimensional condition in open-channel hydraulics where the water depth and velocity remain constant along the channel length, occurring at the normal depth defined by the equality of the bed slope and the friction slope. This state assumes no variation in flow properties with distance, making it a baseline for analyzing steady discharges in engineered waterways. The foundational relation for uniform flow is Chezy's equation, formulated by Antoine Chézy in 1768 during the design of the water supply system, expressing the average V as V = C \sqrt{R S}, where C is the Chezy coefficient, R is the hydraulic radius (cross-sectional flow area divided by wetted perimeter), and S is the channel bed slope. An empirical refinement, Manning's equation, introduced by Robert Manning in 1891 in his paper "On the Flow of Water in Open Channels and ," provides a practical form: V = \frac{1}{n} R^{2/3} S^{1/2}, where n is the Manning roughness coefficient, a measure of channel surface resistance. This equation derives from Chezy's by substituting C = \frac{1}{n} R^{1/6}, linking velocity to channel geometry and roughness in a dimensionally consistent manner suitable for natural and artificial channels. Uniform flow conditions are typically realized in long, prismatic channels—those with uniform cross-section and roughness—under constant , where the component of along the bed balances frictional energy losses. In practice, Manning's equation facilitates the design and analysis of such flows, particularly for sizing irrigation channels to convey steady water volumes efficiently while minimizing erosion or sedimentation.

Gradually and rapidly varied flow

In open-channel , non-uniform conditions arise when the depth varies along the due to changes in , roughness, or cross-section, deviating from the constant normal depth associated with uniform . Gradually varied (GVF) represents a category of such non-uniform where the depth changes slowly over a relatively long distance, allowing inertial forces to dominate while friction and bed effects accumulate gradually. This occurs in prismatic channels under steady conditions with negligible vertical , enabling the use of the energy equation to derive the governing for depth variation. The fundamental equation for GVF is derived from the balance of energy, expressed as: \frac{dy}{dx} = \frac{S_0 - S_f}{1 - \mathrm{Fr}^2} where y is the flow depth, x is the distance along the channel, S_0 is the bed slope, S_f is the friction slope, and \mathrm{Fr} is the Froude number (\mathrm{Fr} = V / \sqrt{g y}, with V as mean velocity and g as gravitational acceleration). The sign of dy/dx determines whether the depth increases (backwater curve) or decreases (drawdown curve), depending on whether the flow is subcritical (\mathrm{Fr} < 1) or supercritical (\mathrm{Fr} > 1). GVF profiles are classified relative to normal depth y_n (uniform flow depth) and critical depth y_c (depth at \mathrm{Fr} = 1), based on the channel slope type: mild (S_0 < S_c, where S_c is critical slope), steep (S_0 > S_c), critical (S_0 = S_c), horizontal, or adverse. Representative GVF profiles include the curve on a mild slope, where subcritical flow has depth greater than normal (y > y_n > y_c), forming a backwater profile upstream of an obstruction, and the S3 curve on a steep slope, where supercritical flow has depth less than critical (y_c > y_n > y), representing a drawdown profile downstream of a . The full system organizes profiles into zones (1 through 3) for each slope type, as summarized below:
ProfileSlope TypeDepth RelationFlow RegimeDescription
Mildy > y_n > y_cSubcriticalBackwater curve
S3Steepy_c > y_n > ySupercriticalDrawdown curve
Horizontaly > y_cSubcriticalBackwater above critical
M2Mildy_n > y > y_cSubcriticalDrawdown to normal
S1Steepy > y_nSubcriticalBackwater to normal
Rapidly varied flow (RVF) contrasts with GVF by featuring abrupt depth changes over short distances, where vertical accelerations and are significant, and is negligible. The primary example is the , a sudden transition from supercritical to subcritical flow that dissipates excess energy through and eddies. This phenomenon is analyzed using the momentum principle, assuming hydrostatic and a across the jump. The sequent depth ratio for a in a rectangular is given by: \frac{y_2}{y_1} = \frac{1}{2} \left[ \sqrt{1 + 8 \mathrm{Fr}_1^2} - 1 \right] where y_1 and y_2 are the upstream and downstream depths, respectively, and \mathrm{Fr}_1 is the upstream . Energy loss in the jump, \Delta E = \frac{(y_2 - y_1)^3}{4 y_1 y_2}, arises from the momentum balance and increases with \mathrm{Fr}_1, making jumps effective for energy dissipation in designs. RVF, including jumps, is commonly caused by sudden obstructions such as weirs, gates, or abrupt transitions that force supercritical flow into subcritical conditions.

Governing Equations

Continuity and mass conservation

In open-channel flow, the principle of ensures the , assuming incompressible behavior typical of under standard conditions. Under the one-dimensional , which averages properties over the cross-section and neglects transverse variations, the volumetric Q remains constant along the for steady , expressed as Q = A V, where A is the wetted cross-sectional area and V is the mean velocity. This equation derives from the integral form of the mass conservation law applied to a fixed in the . For an incompressible with \rho, the general statement is \frac{d}{dt} \int_V \rho \, dV + \int_S \rho \mathbf{v} \cdot d\mathbf{A} = 0, where V is the control volume and S its surface. In steady open-channel flow without lateral inflows, the time derivative vanishes, and assuming no mass flux through the free surface or , the net flux simplifies to zero discharge difference between upstream and downstream sections, yielding \frac{dQ}{dx} = 0 or constant Q. For channels with varying width, this form accounts for changes in A due to geometry, as A incorporates the local width in its definition. For unsteady flow, the extends to \frac{\partial A}{\partial t} + \frac{\partial Q}{\partial x} = 0, capturing temporal changes in storage within the . This relates variations in flow depth, cross-sectional area, and velocity, enabling predictions of how adjusts with changing conditions or upstream influences. In practice, it underpins calculations for flow capacity in natural rivers or engineered channels, where depth measurements inform area and thus velocity estimates via V = Q / A.

Momentum principles

The momentum principles in open-channel flow are derived from the Navier-Stokes equations by integrating over the flow depth and applying simplifying assumptions suitable for shallow water conditions, where the horizontal length scale significantly exceeds the vertical depth. Key assumptions include hydrostatic pressure distribution, negligible vertical accelerations, and a horizontally velocity profile across the cross-section. This depth-integration process yields the one-dimensional Saint-Venant equations, which couple the with the equation to describe unsteady flow dynamics. The equation for open-channel flow expresses the conservation of along the channel axis, accounting for temporal changes in , convective , forces, gravitational components, and frictional : \frac{\partial Q}{\partial t} + \frac{\partial}{\partial x} \left( \alpha \frac{Q^2}{A} \right) + g A \left( \frac{\partial h}{\partial x} + S_f \right) = 0 Here, Q is the , t is time, x is the streamwise distance, A is the cross-sectional flow area, g is , h is the piezometric head ( surface elevation relative to a datum), S_f is the friction , and \alpha is the correction factor that adjusts for non-uniform distributions across the section (typically \alpha \approx 1.0 to $1.2 for turbulent flows in wide channels). The term g A \frac{\partial h}{\partial x} represents the net hydrostatic due to the surface , while the gravity component is incorporated via the bed in \frac{\partial h}{\partial x} = \frac{\partial y}{\partial x} - S_0, where y is flow depth and S_0 is bed . Frictional effects are captured by S_f, often derived from empirical laws. These equations, known as the Saint-Venant equations, are hyperbolic partial differential equations solved numerically for unsteady flows. Hydrodynamic forces in open-channel momentum analysis primarily consist of pressure forces (hydrostatic, assuming linear distribution with depth) and the component of parallel to the bed slope, balanced against flux and boundary shear stresses. The hydrostatic assumption simplifies the pressure term to \int_0^y \rho g (y - z) \, dz per unit width, integrated over the wetted area. For unsteady flows, the Saint-Venant momentum equation models propagation, where a sudden change in (e.g., from gate operation) generates a moving hydraulic discontinuity that travels at a celerity influenced by flow depth and , typically analyzed via characteristics methods. A key application is the , an abrupt transition from supercritical to subcritical flow where balance determines the sequent depths. Applying the equation to a across the jump in a rectangular , the net hydrostatic \frac{1}{2} \rho g (h_2^2 - h_1^2) b (with width b, upstream depth h_1, downstream depth h_2) equals the change in flux \rho Q (V_2 - V_1), neglecting friction over the short jump length. With Q = h_1 V_1 = h_2 V_2, this yields the sequent depth relation: \frac{h_2}{h_1} = \frac{1}{2} \left( -1 + \sqrt{1 + 8 \text{Fr}_1^2} \right) where \text{Fr}_1 = V_1 / \sqrt{g h_1} is the upstream Froude number. This balance highlights momentum conservation in energy-dissipating structures like spillways.

Energy considerations

In open-channel flow, the energy equation describes the conservation of mechanical energy along a streamline, adapted from the Bernoulli equation to account for the free surface. The total head H at any section is given by H = z + y + \frac{V^2}{2g}, where z is the bed elevation above a datum, y is the flow depth, V is the mean velocity, and g is gravitational acceleration. Between two sections along the channel, this total head remains constant in the absence of losses, but friction and other dissipations reduce it downstream: H_1 = H_2 + h_L, where h_L represents the head loss. Specific energy E, which ignores bed elevation changes and focuses on energy relative to the local channel bottom, is defined as E = y + \frac{V^2}{2g}. For a given discharge Q, plotting E versus y yields the specific energy diagram, a curve that exhibits a minimum value at the critical depth where the Froude number Fr = 1. This minimum occurs because, for rectangular channels, E reaches its lowest point when y_c = \left( \frac{q^2}{g} \right)^{1/3}, with q = Q/b as the discharge per unit width and b the channel width; subcritical depths ( y > y_c, Fr < 1) yield higher E, while supercritical depths ( y < y_c, Fr > 1) do the same on the other branch. These diagrams illustrate how flow can transition between alternate depths for the same specific energy and discharge, aiding analysis of controls like weirs or bumps. Energy losses in open-channel flow primarily arise from friction along the wetted perimeter and minor effects at transitions. Frictional losses are commonly quantified using , where the energy slope S_f = \frac{n^2 V^2}{R_h^{4/3}} and head loss h_f = S_f L, with n as the , R_h the , and L the ; values of n range from 0.012 for smooth to 0.035 for natural streams. Alternatively, the Darcy-Weisbach equation applies for more precise turbulent flow modeling: h_f = f \frac{L V^2}{2 g D_h}, where f is the (dependent on and relative roughness) and D_h = 4 R_h the . Minor losses, such as those from sudden expansions or contractions, are typically smaller and expressed as h_m = K \frac{V^2}{2g}, with K an empirical ; these are gradual in slowly varying flows but can be significant in abrupt transitions. The Bernoulli principle, adapted for the by treating hydrostatic pressure up to y, thus reveals pairs of alternate depths (sub- and super-critical) that convey the same Q at equal E, enabling design of energy-efficient channels.

Analysis and Applications

Flow computation methods

Flow computation methods for open-channel flow primarily address gradually varied flow (GVF) conditions, where water surface profiles deviate from uniform flow due to changes in channel geometry, slope, or boundary controls. These methods solve the governing and equations numerically to determine depth, , and along the channel. Practical techniques range from manual step-by-step calculations to advanced software simulations, enabling engineers to predict flood levels, design structures, and assess . The direct step method is an explicit numerical approach for computing GVF profiles, particularly useful for prismatic channels where depth changes are integrated along the channel length. It involves selecting increments in water depth (Δy) and calculating the corresponding distance (Δx) over which the change occurs, effectively integrating the differential equation dy/dx from the gradually varied flow profile equation over discrete segments. Starting from a known depth at a control section (such as critical depth or a downstream boundary), the method computes specific energy, friction slope, and bed slope at each step to determine Δx = ΔE / (S₀ - S̄), where ΔE is the energy difference, S₀ is the bed slope, and S̄ is the average friction slope; this process repeats upstream or downstream until the desired profile is obtained. The method assumes hydrostatic pressure and prismatic geometry, making it efficient for hand calculations or simple programming without iteration. In contrast, the standard step method balances or between consecutive cross-sections separated by a fixed distance (Δx), solving for the unknown depth at the next section through . This technique applies the , incorporating conveyance and losses, to step progressively upstream in subcritical or downstream in supercritical , often using trial-and-error to match the grade line. It is particularly suited for irregular channels with varying , as implemented in software that handles multiple cross-sections defined by surveyed data. balancing may be used alternatively when methods fail near hydraulic jumps or strong accelerations, ensuring principles from the governing equations are maintained. For more complex scenarios involving unsteady flow or extensive river systems, numerical models solve the one-dimensional Saint-Venant equations using schemes to simulate temporal and spatial variations in flow depth and discharge. These models discretize the channel into computational elements, applying implicit or explicit schemes to approximate the hyperbolic partial differential equations while incorporating boundary conditions like inflows or reservoirs. The software, developed by the U.S. Army Corps of Engineers, exemplifies this approach for one-dimensional unsteady flow analysis, supporting mixed flow regimes and integrating with GIS data for real-world applications such as flood routing. Backwater computations focus on subcritical flow regimes, where downstream controls propagate effects upstream, requiring profile calculations from known boundary conditions like reservoir tailwater levels. In such cases, the standard or direct step methods proceed upstream from the control point, adjusting depths to account for the rising water surface due to obstructions or storage. For example, near a reservoir, the backwater curve (M1 profile) extends upstream until normal depth is reached, influencing flood elevations over long distances in mild slopes. These methods rely on key assumptions, including hydrostatic pressure distribution, which holds only for gradually varied conditions without significant vertical accelerations, and one-dimensional flow averaging across the cross-section. Limitations arise in rapidly varied flows, steep slopes exceeding 10%, or non-prismatic channels where analytical solutions like the Manning equation suffice for uniform flow but numerical approaches may require calibration against field data. Computational efficiency decreases with finer grids in numerical models, and errors can accumulate from approximations, necessitating validation for high-impact applications.

Hydraulic design principles

Hydraulic design principles for open-channel flow emphasize , , and to convey water while minimizing , , and environmental disruption. Key criteria include selecting permissible velocities that prevent scour or deposition based on channel material. For non-cohesive materials like fine sand, velocities should not exceed 2.0 ft/s (0.61 m/s), while coarser materials such as clay can tolerate up to 6.0 ft/s (1.83 m/s); these limits are derived from analyses incorporating and flow impingement. In cohesive soils, permissible velocities range from 2.0 ft/s for sandy to 3.5 ft/s for clay, with adjustments for vegetation like Bermuda grass increasing tolerance to 6.0 ft/s. Manning's roughness coefficient () is selected according to lining to account for in calculations; for smooth , ≈ 0.013–0.015, whereas channels with clean, straight banks use = 0.035. Channel sizing prioritizes sections that maximize for a given cross-sectional area and slope, often using uniform flow concepts to determine normal depth. The semicircular section is theoretically the most efficient, as it minimizes wetted perimeter and maximizes hydraulic (R_H = r/2, where r is ), discharging more water than other shapes under identical conditions. For earth-lined channels, trapezoidal sections are practical and efficient, typically with side slopes of 1:1 to 2:1 (horizontal:vertical) to ensure bank stability while approximating the semicircle's optimality; for the most hydraulically efficient trapezoidal section, the side slopes are z:1 with z = 1/√3 ≈ 0.577 (angle of approximately ° to horizontal), and the bottom width b = (2/√3) y ≈ 1.155 y, where y is the flow depth, resulting in a hydraulic R_h = y/2. Hydraulic controls such as weirs and chutes are integral for managing energy in varied flow regimes, particularly to dissipate excess energy and control water surface profiles along the grade line. Weirs in drop structures, like straight or box inlet types, create vertical drops (0.6–15 times critical depth) to form hydraulic jumps, with basin lengths designed as L_B = 2.55 y_c + nappe strike distance (y_c = critical depth) and tailwater depths at least 2.15 y_c to contain the jump. Chutes handle supercritical flows in steep sections, using expansions or roughness elements (e.g., baffle blocks spaced at 0.75–1.0 times incoming depth) to transition to subcritical flow; for example, in USBR Type III basins, end sill heights are set per empirical charts to optimize energy loss. Grade lines are established to align energy gradients with terrain, ensuring controlled varied flow without excessive scour. Modern designs incorporate environmental and climate considerations to enhance . For fish passage, channels must maintain hydraulic geometry up to bankfull stage, with widths exceeding natural bankfull to allow and riparian ; post-2000 guidelines recommend dynamic for 50–100-year to support aquatic habitats. effects, such as increased magnitudes from intense , necessitate adjustments in design standards; European guidelines post-2007 (e.g., EU Floods Directive) mandate non-stationary risk assessments with peak flow uplifts up to 105% by 2080, updating periods from 1-in-10 to 1-in-50 years for critical channels. Case studies illustrate these principles in practice. On the , levees spanning over 4,600 km from to were hydraulically designed under the Mississippi River and Tributaries Project to contain project design floods, but modeling shows they amplify peak discharges by up to 25% by reducing floodplain storage and accelerating flow. In urban settings, the Thornton Creek Water Quality Channel in retrofitted a stormwater system with a tiered meeting hydraulic residence times, reducing peak flows and improving water quality through vegetated swales and modified cross-sections.

References

  1. [1]
    Chapter 5 Flow in open channels | Hydraulics and Water Resources
    Open channel flow is when water is exposed to the atmosphere, not under pressure. Flow can be uniform, gradually varied, subcritical, supercritical, or ...Missing: fundamentals | Show results with:fundamentals
  2. [2]
    [PDF] CHAPTER 5 OPEN-CHANNEL FLOW - MIT OpenCourseWare
    INTRODUCTION. 1 Open-channel flows are those that are not entirely included within rigid. boundaries; a part of the flow is in contract with nothing at all, ...Missing: fundamentals | Show results with:fundamentals
  3. [3]
    [PDF] Chapter 13 OPEN-CHANNEL FLOW - BYU Engineering
    Open-channel flow is driven by gravity, reaching uniform flow when friction equals elevation drop. Uniform flow has constant depth, while nonuniform flow ...
  4. [4]
    Open Channel Flow
    Open channel flow is defined as fluid flow with a free surface open to the atmosphere. Examples include streams, rivers and culverts not flowing full.
  5. [5]
    Open Channel Flow - an overview | ScienceDirect Topics
    In open channel flow the free surface is always at a constant absolute pressure (usually atmospheric) and the driving force of the fluid motion is gravity. In ...Missing: key | Show results with:key
  6. [6]
    [PDF] Hydraulics 3 Open-Channel Flow: Introduction - 1 Dr David Apsley
    The common feature of all open-channel flows is the free surface, where the gauge pressure 𝑝 = 0. All such flows are gravity-driven, with the discharge 𝑄 and ...Missing: key characteristics
  7. [7]
    Henry Darcy and the making of a law - Brown - 2002 - AGU Journals
    Jul 17, 2002 · A similar equation with different coefficients was used for open channel flow. At high flow velocities, the first order term was often ...
  8. [8]
    [PDF] Robert Manning (A Historical Perspective) - DTIC
    equation for open channel flow has been. ERDCTN-EMRRP-SR-10. Page 4. advanced that has displaced the Manning equation for practicing hydraulic engineers. This ...
  9. [9]
    [PDF] CHAPTER 4 FLOW IN CHANNELS - MIT OpenCourseWare
    Most closed conduits in engineering applications are either circular or rectangular in cross section. Open-channel flows, on the other hand, are those whose.
  10. [10]
    CIVE 530 - OPEN-CHANNEL HYDRAULICS LECTURE 1
    Open-channel flow has free surface subject to atmospheric pressure. Pipe flow (closed-conduit) flow has only hydraulic pressure. Open-channel flow problems ...
  11. [11]
    [PDF] OPEN-CHANNEL HYDR~AULICS
    The text.is Ol:ganitlec,i into five parts-namely, Basic Principles, Uni- form Flow, Gradually Varied Flow, Rapidly Varied Flow, and Unstea,dy. Flow. ... Chow ...
  12. [12]
    [PDF] BASIC HYDRAULIC PRINCIPLES OF OPEN-CHANNEL FLOW
    Definition sketch of the energy diagram for open-channel flow ... could be used as the definition sketch for flow through a bridge opening; for example ...
  13. [13]
    On the classification of open channel flow regimes, Dr. Victor M. Ponce
    Open channel flow has been classified into the following four regimes: laminar-subcritical, turbulent-subcritical, laminar-supercritical, and turbulent- ...<|control11|><|separator|>
  14. [14]
    [PDF] Section 2F-2 - Open Channel Flow
    Open channel flow involves understanding stream hydraulics, using Manning's equation, and considering uniform and non-uniform flow, with concepts like critical ...Missing: fundamentals | Show results with:fundamentals
  15. [15]
    Flow Types - Texas Department of Transportation
    Subcritical flow occurs when the actual flow depth is higher than critical depth. A Froude Number less than 1.0 indicates subcritical flow. This type of ...<|control11|><|separator|>
  16. [16]
    [PDF] Open Channel Hydraulics I – Uniform Flow - PDH Online
    or turbulent, is a value for Reynold's number (Re). Laminar open channel flow occurs for Re less than about 500 and turbulent flow for Re greater than about ...<|control11|><|separator|>
  17. [17]
    [PDF] Uniform Flow - C
    Flow properties remain constant with respect to distance. ● Uniform flow in open channels is understood to mean. steady uniform flow. ● Uniform flow is ...
  18. [18]
    [PDF] The Manning Equation For Open Channel Flow Calculations
    in 1768 by French physicist and engineer Antoine de Chézy (1718–1798) while designing Paris's water canal system. Chézy discovered a similarity parameter that.
  19. [19]
    Manning, Manning formula, history of the Manning formula, Fadi ...
    This formula saw the light in 1891, in a paper written by him entitled "On the flow of water in open channels and pipes," published in the Transactions of the ...
  20. [20]
    [PDF] Uniform Flow and Flow Resistance
    Feb 19, 2017 · The average velocity of a uniform open-channel flow is proportional to the square root of the product of hydraulic radius (R) and the downslope.
  21. [21]
    [PDF] Chapter 3 Hydraulics - USDA
    In open-channel and pipe flow, the slope, or fall, of the energy grade line for a length of channel or pipe represents the loss of energy due to friction.
  22. [22]
    [PDF] Chapter 2. Derivation of the Equations of Open Channel Flow
    In the following sections, the open channel flow equations based on continuity (Section 2.2), momentum (Section 2.3) and energy (Section 2.4) are derived. Since ...
  23. [23]
    [PDF] The derivation of the Saint–Venant equations - The Bulletin of NCC
    In this paper, the rigorous deduction of the shallow-water equa- tions from the Navier–Stokes equations using the recommendations and methods proposed in [1] is ...
  24. [24]
    [PDF] Block 3 – Open channel flow
    Saint Venant equations in 1D. • continuity (for section without inflow). • Momentum equation from integration of. Navier-Stokes/Reynolds equations over.
  25. [25]
    Energy and momentum in one dimensional open channel flow
    Aug 6, 2025 · The equations of continuity, momentum and energy are derived, incorporating depth-dependent correction factors for velocity distributions ...
  26. [26]
    Channel Flow Basic Concepts, Equations, and Solution Techniques
    The physically-based routing models, such as the kinematic-wave model and the Muskingum-Cunge model use Manning's equation and Manning's roughness coefficients ...
  27. [27]
    Modeling positive surge propagation in open channels using the ...
    In this study, the SGN equations for variable bed profiles are solved for simulating the positive surge waves.
  28. [28]
    [PDF] HYDRAULIC JUMP
    The most common application of the momentum equation in open-channel flow ... • The ratio of sequent depths, that is the flow depths upstream and downstream of.
  29. [29]
    [PDF] hydjump.pdf
    A hydraulic jump is an abrupt change from a shallow high-speed flow to a deep low-speed flow of lower energy. It occurs when either the bed slope and changes to ...
  30. [30]
    [PDF] topic 3 - the energy equation and critical flow
    Mar 24, 2003 · CHANNELS. The Energy Equation and Open Channel Flow. • In our analysis of open channel flow, we typically make the following assumptions:.
  31. [31]
    [PDF] Specific Energy
    Consider a channel where the upstream velocity is. 5.0 m/s and the upstream flow depth is 0.6 m. The flow then passes over a bump 15 cm in height.
  32. [32]
    [PDF] Topic 8: Open Channel Flow
    Open channel flow involves the Manning equation, relating velocity to energy loss, and normal depth, the depth of uniform flow. Natural channels are irregular, ...Missing: fundamentals | Show results with:fundamentals
  33. [33]
    Direct Step Backwater Method - Texas Department of Transportation
    In the Direct Step Method, an increment (or decrement) of water depth (δd) is chosen and the distance over which the depth change occurs is computed. The ...
  34. [34]
    Standard Step Backwater Method
    This procedure applies to most open channel flow, including streams having an irregular channel with the cross section consisting of a main channel and separate ...
  35. [35]
    Backwater in Subcritical Flow - Texas Department of Transportation
    Backwater in Subcritical Flow. In subcritical flow conditions, the backwater tails off upstream until it reaches the normal water surface.
  36. [36]
    [PDF] Gradually varied flow (GVF) - C
    Water surface profile computation. ○ Two types of methods. 1. Explicit or direct step method: distance is determined for a specified depth change. •. Mostly ...
  37. [37]
  38. [38]
    [PDF] Water Surface Profile Calculations with HEC‐RAS
    • Basic Open Channel Flow Concepts. • Energy Principles. • Cross Section ... The assumption of a hydrostatic pressure distribution is only valid for slopes less ...
  39. [39]
    [PDF] Hydraulic Design of Flood Control Channels - USACE Publications
    Jun 30, 1994 · This manual presents procedures for the design analysis and criteria of design for improved channels that carry rapid and/or tranquil flows.
  40. [40]
    [PDF] Guide for Selecting Manning's Roughness Coefficients for Natural ...
    CHANNEL n VALUES. The most important factors that affect the selection of channel n values are (1) the type and size of the materials that compose the bed ...
  41. [41]
    [PDF] 2.5 MOST EFFICIENT CROSS SECTION • For steady uniform flow ...
    Feb 16, 2003 · Semi-circular open channel will discharge more water than any other shape (assuming that the area, slope and surface roughness are the same).
  42. [42]
  43. [43]
    Culvert Design Guidelines for Ecological Function
    The purpose of this document is to provide basic culvert design guidelines preferred by the US Fish and Wildlife Service (USFWS) Alaska Fish Passage Program.
  44. [44]
    Climate change adjustments in engineering design standards
    Oct 28, 2021 · Climate change adjustments in engineering design standards, such as design precipitation and design floods, are reviewed via examples from Europe.
  45. [45]
    Assessing the impacts of dams and levees on the hydrologic record ...
    Jul 15, 2018 · To assess the impacts of dams and levees on Mississippi River discharges we used hydrologic analyses, statistical analyses, and hydraulic ...
  46. [46]
    Thornton Creek Water Quality Channel
    The project uses a tiered system to treat stormwater flows. · A modified water quality channel design was developed to meet hydraulic flow and residence time ...