Hydraulic Jump

Keith bechtol october 31, 2007, (submitted as coursework for physics 210, stanford university, fall 2007), introduction.

A hydraulic jump is a fluid shockwave created at the transition between laminar and turbulent flow. One common example of a hydraulic jump can be seen in the water radiating outward when the stream of tap water strikes the horizontal surface of a sink. The water initially flows in a smooth sheet with consistent current patterns. In this region, the speed of the water exceeds the local wave speed. Friction against the sink surface slows the flow until an abrupt change occurs. At this point, the depth increases as water piles up in the transition region and flow becomes turbulent [1]. The motion of individual water molecules becomes erratic and unpredictable. The interruption of flow patterns also reduces the kinetic energy of the water. In addition to the kitchen sink example, hydraulic jumps are also typical features of river rapids where the water swirls and foams around rocks and logs.

Basic Theory

Although the hydraulic jump effect is common to everyday experience and has been studied experimentally for many years, the underlying theory describing the phenomena is surprisingly complex. Rayleigh first described the problem in 1914 [2]. He calculated of the change in fluid depth associated with the shockwave and introduced the concepts of continuity and fluid momentum conservation in his derivation. A schematic of fluid flow at a hydraulic jump is depicted in Figure 1.

Diagram of classical hydraulic jump.

Let region 1 represent section of fast laminar flow preceding the hydraulic jump and region 2 denote the section of slow turbulent flow after the transition. In the present analysis, assume a uniform hydrostatic pressure distribution and a uniform velocity distribution. Consider vertical slices of fluid representing unit areas of the flow. Continuity of fluid flow implies that that the discharge, q=hv , must be equal before and after the hydraulic jump. h denotes the water depth and v is the water velocity.

A useful quantity to define is the Froude number F R =v/(qh) 1/2 . The Froude number is the ratio of flow velocity to wave celerity and marks the boundary between critical and subcritical flow [1]. For conditions in which the Froude number is greater than one, flow velocity exceedes the wave celerity (the speed of an individual wave crest) and the fluid motion is smooth. Subcritical currents with Froude numbers less than exhibit turbulent flow.

Next, call the ratio of fluid depths D=h 1 /h 2 . Solving the two equations above yields the following differential equation in terms of the ratio of fluid depths and Froude number:

The solution of this differential equation provides the ratio of final depth following a hydraulic jump to the initial depth before the jump.

Another important property of the hydraulic jump is the energy dissapated by the transition to turbulent flow. The energy loss is usually measured in terms of change in hydraulic head, H=h+(q 2 /2gA 2 ) where A is the cross sectional area of the flow. Efficiency is written as η=ΔH/H 1 .

Both the change in fluid depth and the energy dissipated can be quantified in terms of the Froude number. As the Froude number increases, the change in fluid depth grows and the energy dissipated by the jump rises.

Diagram of radial hydraulic jump

Radial Hydraulic Jump

The theoretical framework developed by Rayleigh can be also used to predict the location of a radial hydraulic jump. Consider the arrangement in Figure 2 for a radial hydraulic jump. Following Rayleigh's theoretical approach, one can derive the position of a radial hydraulic jump as

Here, R j is the radial position of the jump, d represents the post-jump depth, a is the jet radius, g is gravitational acceleration, and q is jet discharge. Notice that fluid viscosity does not explicitly enter into Rayleigh's calculation of the jump position in this formula. Indeed, the effect of fluid viscosity on shockwave formation is one of the outstanding theoretical questions concerning hydraulic jumps. In the 1960s, Watson developed a new model to account for the change to turbulent flow following the hydraulic jump [3]. Using a series of approximation techniques, Watson proposed

Radial Jump for Superfluid Helium

Contemporary experiments have put fluid dynamics models to an extreme test by measuring the radii of hydraulic jumps using liquid helium. Above the lambda point critical temperature, liquid helium behaves as a normal fluid with conventional viscosity properties. However, as the temperature drops below 2.17 K, the liquid helium experiences a phase change and becomes a superfluid with effectively zero viscosity. The vast majority of hydraulic jump models consider conventional fluids and until recently, the ability of these models to describe superfluids was untested. The importance of the hydraulic jumps in the study of fluid dynamics motivates further investigation into the superfluid scenario.

In 2007, Rolley, Guthmann, and Pettersen performed an experiment to analyze radial hydraulic jumps in liquid helium [4]. A jet of liquid helium falls vertically downward onto the surface of a horizontal mirror. The depth of the liquid helium is measured by a CCD camera positioned at a shallow viewing angle. Additional mirrors are positioned at angles to observe the radial position of the jump from above. Next the jet of liquid helium is gradually cooled from an initial temperature of 4.2 K through the superfluid transition point to a final temperature of 1.5 K. The jump radius is measured at liquid helium temperatures above and below the superfluid transition to gauge the effect of fluid viscosity.

Above the critical temperature, the position of the hydraulic jump position is well described by theory as demonstrated by a close correspondence with experimental values. Models not accounting for surface tension tend to overestimate jump radii while models incorporating surface tension tend to predict jump radii slightly too low. More surprisingly though, the model proposed by Watson continued to accurately predict shockwave conditions of liquid helium below 2.17 K. Rolley, Guthmann, and Pettersen explain this effect by observing that the liquid helium below 2.17 K is actually a mixture of normal and superfluid components [4]. Additionally, superfluidity is disrupted above a critical fluid velocity. Therefore, even at a temperature of 1.5 K, the effective viscosity of the liquid helium was certainly non-zero. However, the experimenters were able to confirm the presence of a superfluid component by observing ripple patterns uncharacteristic of normal fluids.

Despite experimental limitations, the accuracy of theoretical predictions of shockwave conditions over a wide range in fluid viscosities indicates that current models may be applicable to a larger range of fluid behavior than previously expected.

Industrial Application as Energy Dissipator

Hydraulic jumps remain a topic of continued scientific and technological interest in part due to their industrial utility as energy dissipators. One of the most important applications of the hydraulic jump is to reduce the impact of dams downstream. Rapid outflow from a spillway erodes the channel and can undermine the structural strength of the reservoir if left unchecked. However, hydraulic jump stilling basins can reduce the discharge energy by up seventy percent [5]. Internal friction and mixing high velocity flow into the larger water volume lowers the speed of outflow. Depending on the type of design, changes in slope, channel width, or obstacles positioned along the spillway trigger the transition to turbulent flow. Forces from hydrostatic uplift, cavitation, vibration, and abrasion create unique engineering challenges for each class of spillway. Consequently, the design of a practical energy dissipator must balance efficiency with durability.

Termination Shock Analog

The termination shock of the Solar System is a more exotic analog of the hydraulic jump. The Sun emits a flow of charged particles traveling outward at speeds of 420 km/s. Along the way, plasma interactions with the interstellar medium slow the particles. The point at which the solar wind drops to a subsonic speed of about 100 km/s in the interstellar medium is called the termination shock. In this region, the solar wind experiences compressional heating, pressure fluctuations, and a sudden magnetic field changes. Beyond the termination shock, the solar wind is effectively stopped by the interstellar medium in a region called the heliosphere. The bow shock enveloping the heliosphere marks the outer edge of the Solar System.

The outer reaches of the Solar System remained largely unexplored until the last decade. Two Voyager Mission spacecraft launched in 1977 offered the best hope of directly probing this region. By 1993, Belcher, Lazarus, McNutt, and Gordon were able to use early observations of the Voyager 1 spacecraft to predict the location of the termination shock [6]. Then in December 2004, Voyager 1 entered a region with high intensities of low-energy (~1 MeV) solar wind particles at a distance of 94 AU from the Sun [7]. The encounter marked the first direct observation of the termination shock. This discovery required the efforts of both Voyager spacecraft simultaneously exploring separate regions of the Solar System. Since the solar wind pressure varies over time due to changing solar activity levels, Voyager 2 was needed as a calibration instrument to distinguish the termination shock from pressure fluctions due to normal solar events. Voyager 2 lags Voyager 1 by about 20 AU in its outward trajectory and the two spacecraft travel in different directions to probe distinct regions of the Solar System boundary. Subsequent observations have revealed that the position of the termination shock shifts by about 10 AU during an eleven-year solar acitivity cycle. Although the plasma interactions involved in the termination shock differ substantially from hydraulic jumps in water, the abrupt transition of fluid behavoir when dropping to subcritical flow velocities is shared by both phenomena.

© 2007 Keith Bechtol. The author grants permission to copy, distribute and display this work in unaltered form, with attributation to the author, for noncommercial purposes only. All other rights, including commercial rights, are reserved to the author.

[1] W. H. Hager, Energy Dissipators and Hydraulic Jump , (Kluwer Academic Publishers, 1992).

[2] L. Rayleigh, "On the Theory of Long Waves and Bores," Proc. Roy. Soc. Lond. A 90 , 324 (1914).

[3] E. J. Watson, "The spread of a Liquid Jet Over a Horizontal Plane", J. Fluid Mech. 20 , 481 (1964).

[4] E. Rolley, C. Guthmann and M. S. Pettersen, "The Hydraulic Jump and Ripples in Liquid Helium", Physica B: Cond. Mat. 394 , 46 (2007).

[5] R. M. Khatsuria, Hydraulics of Spillways and Energy Dissipators , (Marcel Decker, 2005).

[6] J. W. Belcher, A. J. Lazarus, R. L. McNutt Jr., and G. S. Gordon, "Solar Wind Conditions in the Outer Heliosphere and the Distance to the Termination Shock", J. Geophys. Res. 98 , 177 (1993).

[7] W. R. Webber, "An Empirical Estimate of the Heliospheric Termination Shock Location with Time with Application to the Intensity Increases of MeV Protons Seen at Voyager 1 in 2002-2005", J. Geophys. Res. 110 , 209 (2005).

froude experiment

Fluid flow: Froude and Reynolds numbers

  • November 12, 2021
  • No Comments

froude experiment

19 th century experiments that helped quantify the nature of fluid flow, surface waves, and bedforms .

It all depends on inertia , like the reluctance to get out of bed on a cold winter’s morning. But rather than feeling guilty, acknowledge that by sleeping in you are adhering to the mechanical Laws that prevent the universe from collapsing – the inertial forces that keep planets in orbit around their suns, and suns in motion through their galaxies.

Inertia is loosely defined as a force that resists the change in motion of a body; here motion refers to a vector that describes velocity and direction, and ‘body’ refers to pretty well anything composed of matter, including a body of fluid. The term was coined by astronomer Johannes Kepler (17 th century); his erstwhile colleague Galileo demonstrated its qualities by experimenting with balls rolling along sloping surfaces.

However, it was Isaac Newton who codified the properties of inertia in his three Laws of Motion – apparently Newton credits Galileo with the discovery. The 1 st Law, also called the Law of Inertia , states that the motion of a body will not change unless an external force acts on it (i.e., to accelerate, decelerate, or change its direction). The 2 nd Law quantifies the relationship between an external force F, mass (m) and acceleration (a) as F = ma . And the 3 rd Law states that when an external force is applied, there will be an equal and opposite force that resists the change in motion, i.e., an inertial force – also called the Action-Reaction Law .

Inertial forces depend on the mass of a body – the larger the mass, the greater the force. In fact, the concept of mass itself is based on inertia.

Inertia and fluid flow

Inertial forces are central to the quantification of fluid mechanics. We have William Froude (1810-1879) and Osborne Reynolds (1842-1912) to thank for their eponymous numbers ( Froude number and Reynolds number ) that describe the characteristic states of flow. And because these numbers are dimensionless, they allow experiments with models (e.g., wind tunnels, sediment flumes) that can be scaled to real-world fluid flow phenomena. Scaling can be applied to almost anything related to fluid flow – from the motion of a boat through water, to quantifying the formation of sediment bedforms or sediment gravity flows from small-scale sediment flume experiments. The importance of Froude and Reynolds numbers cannot be overstated.

Froude number

Froude’s influential paper of 1861 was published by the Institute of Naval Architects (PDF available). Froude had surmised that, to predict the behaviour of a ship moving through water, he would need to experiment with much smaller versions of ships, or models , that could be scaled to the behaviour of much larger vessels. Thus, Froude’s number was derived from experiments with model boats, a few metres long.

The number expresses the characteristics of flow, including surface waves and bedforms, as the ratio between inertial forces and gravitational forces:

                                                          Fr = V/√(g.D)

Where V is bulk flow velocity (having dimensional units L.T -1 ) that reflects the dominant effect of inertia on surface flows, and the component √(g.D) where g is the gravitational constant (units of L.T -2 ), and D is water depth (units of L). The denominator represents the speed of a surface gravity wave relative to the bulk flow velocity (√(g.D) simplifies to units of velocity). Whether the surface wave is faster, slower or the same speed as the bulk flow will depend on its resistance to move, or its inertia. Fr is dimensionless.

The numerical value of Fr is used to define three conditions of flow. If Fr = 1 (numerator = denominator), then any surface wave will remain stationary – it will not move upstream or downstream. This condition occurs when both the velocities and water depth are at critical values. Not surprisingly, this condition is called critical flow . A common manifestation of critical flow is the formation of stationary waves (or standing waves) above and usually in phase with antidune bedforms (i.e., upper flow regime ).

A plot of the experimentally determined stability fields for bedforms, as a function of grain size and flow velocity. The transitions from one field to another are abrupt or gradual as indicated. Modified from Ashley, 1990, Figure 1 with minor additions.

When Fr < 1, inertial forces dominate, and the result is a subcritical condition – tranquil flow . This corresponds to lower flow regime bedforms such as ripples and larger dune structures.

When Fr > 1, gravitational forces dominate resulting in supercritical flow conditions. The corresponding stream flow surface conditions manifest as an acceleration of flow such that stationary waves break upstream (chutes – upper flow regime), commonly followed by a rapid decrease in flow and formation of a hydraulic jump where Fr < 1 ( chute and pool conditions). A hydraulic jump is the region of turbulence that represents the transition from supercritical (laminar) flow to tranquil flow – as shown in the kitchen sink example below. Supercritical flow is also common in pyroclastic density currents .

A kitchen sink demonstration of the transition from laminar, supercritical flow to turbulent subcritical flow via a hydraulic jump.

The complexity of flow transitions in a small natural system is shown in this video clip of supercritical and subcritical (tranquil) domains in a small, shallow stream. The standing waves (left) represent critical conditions where the speed of the waves matches the stream flow velocity. Supercritical conditions downstream produce chutes. Downstream migrating ripples in the foreground indicate subcritical flow.

Reynolds number

Schematic representation of laminar and turbulent flow using hypothetical flow lines. The blue arrow (right) indicates mean flow velocity for turbulent flow.

Unlike Froude who was more concerned with the surface configurations of a flowing medium, Reynolds experiments in glass pipes were concerned with the bulk structure of flow, in particular the transition from laminar to turbulent flow ( Reynolds, 1883 , PDF available). To picture this, think of a flowing fluid as a set of flow lines. In laminar flow, the flow lines are parallel, or approximately so, and relatively straight. The flow velocity will be the same across each flow line. By contrast, turbulence is described by flow lines that constantly change direction and velocity. In a flowing stream this is manifested as eddies, boils, and breaking waves. In sedimentary systems, turbulence is an erosive process, and an important mechanism for maintenance of sediment suspension through water columns and in sediment gravity flows.

The video below shows the abrupt transition from laminar flow in the slightly sinuous trail of smoke, to turbulent flow above.

To understand the nature of the laminar-turbulent flow transition, Reynolds considered four variables:

  • Fluid density ρ (units of M.L -3 ).
  • Fluid viscosity ( μ ) that measures the resistance to shear and is strongly temperature-dependent. μ has units of M.(L.T) -1
  • Mean velocity of flow V , that reflects shear rate and inertia forces (units of L.T -1 ), and
  • Tube diameter D that influences the degree of turbulence (units of L).

Reynold’s number is written as:

                                                                  Re = ρVD/μ

that expresses the ratio of inertial (resistance) forces to viscous (resistance) forces . Re is dimensionless.

In his glass tube experiments, Reynolds systematically varied μ , V , and D ( μ was varied by heating the water). For each combination he discovered that the transition from laminar to turbulent flow in water was abrupt, and consistently had Re values of about 12000. Reversing the experiment gave values of about 2000 for the transition from turbulent to laminar flow.

Reynolds’ original glassware used in his fluid flow experiments. Tube diameters ranged from 2.54 cm to 0.62 cm. Coloured dye was introduced through a funnel. In all experimental runs, the transition from laminar to turbulent flow was abrupt. These figures are from Reynolds’ 1863 paper.

Re can be used to determine the kind of flow in large and small fluid systems. As a general rule:

  • Re values <2000 indicate laminar flow ,
  • Re >4000 turbulent flow , and
  • the region in between these two extremes reflects transitional flow .

Flow in most open-surface geological and geomorphic systems tends to be turbulent, with familiar examples including channelized flow (river, tidal and submarine channels) and more open flow across broad expanses such as continental shelves. It also includes volcaniclastic systems like pyroclastic flows and surges. Experimental flow in flumes produces a variety of bedforms at Re values that range from about 4000 to >100,000.

Laminar flow at low velocities is probably responsible for deposition of lower flow-regime plane beds; Allen (1992) has suggested that laminar flow at higher velocities may be restricted to thin sheet floods. Fluids having high viscosity, such as glacial ice and lava, commonly exhibit laminar flow. The Re value in microscopic rock-fluid systems, such as intercrystal boundaries in diagenetic environments, will also be low because fluid viscosity will dominate in such confined spaces.

Comparing Froude and Reynolds numbers

Froude numbers express a relationship between the free-surface of a flow and the various waves and ruffles that form there, and bedforms at the sediment-water interface. Reynolds numbers deal to the bulk characteristics of flow – whether it has laminar or turbulent structure.

The numbers Fr and Re are like chalk and cheese – they are not comparable. Both are dimensionless ratios, but that’s where the similarity ends. Both functions depend on inertial forces (the resistance to do anything), but for Fr the inertial component is in the denominator, and for Re in the numerator. Thus, if inertial forces become dominant, the numerical value of Fr decreases and that for Re increases.

Both numbers have application well beyond the relatively narrow field of sedimentology. Both are used extensively in scaled models – Fr for elucidating the efficacy of movement through a fluid – boats through water, airplanes through air. Re is used extensively to describe fluid flow in biological systems.

Allen (1992) has given sedimentologists a diagram that generalizes the relationship between Fr and Re in terms of mean flow velocity and flow depth. The boundaries of the 4 domains correspond to critical flow transitions; subcritical (tranquil) to supercritical for Fr , and laminar to turbulent for Re . I have added the most common bedforms to these domains.

J.R.L. Allen’s (modified slightly from 1992, Fig. 1.21) plot showing four domains of fluid flow, the boundaries of which are defined by the laminar-turbulent flow transition (Reynolds), and the subcritical-supercritical flow transition (Froude).

There are many publications on this topic, but I highly recommend two publications that provide greater detail of theory and practice on this and other topics in fluid flow and sedimentation:

John Southard’s excellent (open access), online Introduction to Fluid Motion and Sediment Transport .

J.R.L. Allen 1992 (and later editions) Principles of Physical Sedimentology (that no sedimentologist should be without).

Other posts in this series

Identifying paleocurrent indicators

Measuring and representing paleocurrents

Crossbedding – some common terminology

Sediment transport: Bedload and suspension load

The hydraulics of sedimentation: Flow regime

Fluid flow: Shields and Hjulström diagrams

Fluid flow: Stokes Law and particle settling

Leave a Comment Cancel Reply

Your email address will not be published. Required fields are marked *

Save my name, email, and website in this browser for the next time I comment.

Search

Copyright © 2024 Geological Digressions 

  • How membership works
  • Membership Benefits

Medals, Prizes & Awards

Student Ambassador

Scholarships

  • RINA Scholarships
  • Sir William White Scholarship
  • William Froude Scholarship

Branch finder...

Professional Membership

  • Associates (AssocRINA)
  • Graduate Associate Member (AMRINA)
  • Associate Member (AMRINA)
  • Member (MRINA)
  • Fellow (FRINA)

Membership Fees

Professional Registration

  • Registered Professional Engineer Queensland (RPEQ)

Continuing Professional Development (CPD)

William Froude: the father of hydrodynamics

by RINA Editorial | 20th January 2024 | Maritime History & Heritage , RINA News

froude experiment

Froude's combination of calculations and model-based experimentation became the blueprint for measuring water resistance

RINA Historian Mark Barton looks at the life and career of one of the Institution’s, and naval architecture’s, most influential figures

William Froude (1810-1879) was the first person to formulate reliable laws for the resistance that water offers to ships (such as the hull speed equation) and thus enable ship designers to predict their stability and performance. Having worked on railways, Froude was invited by Brunel to turn his attention to the stability of ships in a seaway. He was an early contributor to the then-Institution of Naval Architects and his paper On the Rolling of Ships , was presented at the Institution’s second session in March 1861. Froude would prove hugely influential, and with his associate Henry Brunel, obtained funds from the Admiralty to identify the most efficient hull shape.

He undertook this research by building scale models and established a formula (now known as the Froude number) by which the results of small-scale tests could be used to predict the behaviour of full-sized hulls. He built a sequence of 3, 6 and 12 foot scale models and used them in towing trials. These took place on the River Dart and enabled him to establish their hull resistance and scaling laws, known as the ‘Law of Comparison’. His experiments were vindicated in full-scale trials conducted by the Admiralty. However, he recognised that exposure to wind and waves was impacting adversely on his results.

Torquay Froude House resized

William Froude’s Torquay home. Source: Mark Barton

As a result, the first ship model towing tank or ‘Ship Tank’ was built, for a sum of £2,000, from the Admiralty. This allowed Froude to build and operate an enclosed experiment facility with a carriage, or ‘truck’ as Froude called it, to tow models at steady speed from one end of the tank to the other whilst measuring their drag under controlled conditions.

1010px William Froude and the Admiraltys First Naval Test Tank at Torquay Devon C 1872 HU82582

Froude’s original test tank circa 1872.

The Admiralty approved the plan but increased the scope of supply to include tests on the rolling of ships and highlighted that any cost overrun would have to be borne by Froude himself. This 85 m long tank was built at his home in Torquay in 1872 and the first model tested being the sloop HMS Greyhound . Here he was able to combine mathematical expertise with practical experimentation to such good effect that his methods are still followed today.

Although Froude died in 1879 at the age of 68, his son Robert Edmund (Eddie) Froude continued the research and established a new experimental facility on a redundant plot adjacent to the naval gunboat yard at Haslar. The last test at Torquay was completed on 5th January 1886 and the new tank at the Admiralty Experiment Works, Haslar opened a month later on 6th February. The new tank was larger and 122 m long.

The new facility was used intensively and tested models of all of the major classes of British warship that fought in World War One. The first submarine models were tested in early 1902, just a couple of months before HMS Holland I commenced her sea trials. Capacity limitations during the early inter-war years led to the Admiralty approving a second, larger tank built at right angles to the first.

The new tank (known as No.2 Ship Tank with the original tank becoming No.1) was 271 m long and had an adjustable false floor to allow experiments to be conducted regarding the impact of shallow water. It remains in use today and has been operated by QinetiQ since 2001.

Related Posts

Wind Propulsion 2024: Advancements in sails

Wind Propulsion 2024: Advancements in sails

RINA News , The Naval Architect - News

By Alberto Llopis Pascual, head of aerodynamics, bound4blue  Since the decline of commercial sail-powered vessels, there has been remarkable progress in sail technology, driven by a renewed interest in sustainable maritime solutions. The modern sail has evolved far...

A perfect storm brewing for wind propulsion era 2.0

A perfect storm brewing for wind propulsion era 2.0

New regulations, price challenges for existing and new fuels along with the growing pressure from cargo owners to reduce Scope 3 emissions are driving an increased deployment of wind propulsion technologies, writes Gavin Allwright In 1860, just before he became...

Maritime industry shines at RINA Annual Dinner

Maritime industry shines at RINA Annual Dinner

Industry News , RINA News

The Royal Institution of Naval Architects (RINA) held its highly anticipated Annual Dinner at the De Vere Grand Connaught Rooms in Holborn for the second consecutive year, drawing an impressive crowd of 370 attendees. The event, marked by distinguished speeches,...

Press release: RINA and MARIN to host Human Factors 2024 Technical Conference

Press release: RINA and MARIN to host Human Factors 2024 Technical Conference

Press Releases , RINA News

The Royal Institution of Naval Architects (RINA) and Maritime Research Institute Netherlands (MARIN) to host the Technical Conference: Human Factors 2024 in Wageningen, Netherlands in October 2024 London – 22 May 2024: On 8-10 October 2024, maritime industry...

Encyclopedia Britannica

  • History & Society
  • Science & Tech
  • Biographies
  • Animals & Nature
  • Geography & Travel
  • Arts & Culture
  • Games & Quizzes
  • On This Day
  • One Good Fact
  • New Articles
  • Lifestyles & Social Issues
  • Philosophy & Religion
  • Politics, Law & Government
  • World History
  • Health & Medicine
  • Browse Biographies
  • Birds, Reptiles & Other Vertebrates
  • Bugs, Mollusks & Other Invertebrates
  • Environment
  • Fossils & Geologic Time
  • Entertainment & Pop Culture
  • Sports & Recreation
  • Visual Arts
  • Demystified
  • Image Galleries
  • Infographics
  • Top Questions
  • Britannica Kids
  • Saving Earth
  • Space Next 50
  • Student Center

Above the clouds 130 nautical miles below, astronaut Mark C. Lee floats freely without tethers as he tests the new Simplified Aid for Extravehicular Activity (EVA) Spacewalk Rescue (SAFER) system, Sept. 16, 1994. Space Shuttle Discovery, STS-64

Froude number

Our editors will review what you’ve submitted and determine whether to revise the article.

Froude number (Fr) , in hydrology and fluid mechanics , dimensionless quantity used to indicate the influence of gravity on fluid motion. It is generally expressed as Fr = v /( gd ) 1 / 2 , in which d is depth of flow, g is the gravitational acceleration (equal to the specific weight of the water divided by its density , in fluid mechanics), v is the celerity of a small surface (or gravity) wave, and Fr is the Froude number. When Fr is less than 1, small surface waves can move upstream; when Fr is greater than 1, they will be carried downstream; and when Fr = 1 (said to be the critical Froude number), the velocity of flow is just equal to the velocity of surface waves. The Froude number enters into formulations of the hydraulic jump (rise in water surface elevation) that occurs under certain conditions, and, together with the Reynolds number , it serves to delineate the boundary between laminar and turbulent flow conditions in open channels.

cropped-cropped-Penguins.jpg

Elin Darelius & Team

Elin Darelius & Team's Scientific Adventures

Who is faster, the currents or the waves? The Froude number

froude experiment

A very convenient way to describe a flow system is by looking at its Froude number. The Froude number gives the ratio between the speed a fluid is moving at, and the phase velocity of waves travelling on that fluid. And if we want to represent some real world situation at a smaller scale in a tank, we need to have the same Froude numbers in the same regions of the flow.

For a very strong example of where a Froude number helps you to describe a flow, look at the picture below: We use a hose to fill a tank. The water shoots away from the point of impact, flowing so much faster than waves can travel that the surface there is flat. This means that the Froude number, defined as flow velocity devided by phase velocity, is larger than 1 close to the point of impact.

froude experiment

At some point away from the point of impact, you see the flow changing quite drastically: the water level is a lot higher all of a sudden, and you see waves and other disturbances on it. This is where the phase velocity of waves becomes faster than the flow velocity, so disturbances don’t just get flushed away with the flow, but can actually exist and propagate whichever way they want. That’s where the Froude number changes from larger than 1 to smaller than 1, in what is called a hydraulic jump . This line is marked in red below, where waves are trapped and you see a marked jump in surface height. Do you see how useful the Froude number is to describe the two regimes on either side of the hydraulic jump?

froude experiment

Obviously, this is a very extreme example. But you also see them out in nature everywhere. Can you spot some in the picture below?

froude experiment

But still, all those examples are a little more drastic than what we would imagine is happening in the ocean. But there is one little detail that we didn’t talk about yet: Until now we have looked at Froude numbers and waves at the surface of whatever water we looked at. But the same thing can also happen inside the water, if there is a density stratification and we look at waves on the interface between water of different densities. Waves running on a density interface, however, move much more slowly than those on a free surface. If you are interested, you can have a look at that phenomenon  here . But with waves running a lot slower, it’s easy to imagine that there are places in the ocean where the currents are actually moving faster than the waves on a density interface, isn’t it?

For an example of the explanatory power of the Froude number, you see a tank experiment we did a couple of years ago with Rolf Käse and Martin Vogt ( link ). There is actually a little too much going on in that tank for our purposes right now, but the ridge on the right can be interpreted as, for example, the Greenland-Scotland-Ridge, making the blue reservoir the deep waters of the Nordic Seas, and the blue water spilling over the ridge into the clear water the Denmark Strait Overflow. And in the tank you see that there is a laminar flow directly on top of the ridge and a little way down. And then, all of a sudden, the overflow plume starts mixing with the surrounding water in a turbulent flow. And the point in between those is the hydraulic jump, where the Froude number changes from below 1 to above 1.

froude experiment

Nifty thing, this Froude number, isn’t it? And I hope you’ll start spotting hydraulic jumps every time you do the dishes or wash your hands now! 🙂

All pictures in this post are taken from my blog “ Adventures in Oceanography and Teaching “. Check it out if you like this kind of stuff — I do! 🙂

The Froude number, Fr, is a dimensionless value that describes different flow regimes of open channel flow . The Froude number is a ratio of inertial and gravitational forces.

·          Gravity (numerator) - moves water downhill

·          Inertia (denominator) - reflects its willingness to do so.

V =  Water velocity

D = Hydraulic depth (cross sectional area of flow / top width)

g = Gravity

Fr = 1,     critical flow,

Fr > 1,     supercritical flow (fast rapid flow),

Fr < 1,     subcritical flow (slow / tranquil flow)

The Froude number is a measurement of bulk flow characteristics such as waves, sand bedforms, flow/depth interactions at a cross section or between boulders.

The denominator represents the speed of a small wave on the water surface relative to the speed of the water, called wave celerity. At critical flow celerity equals flow velocity. Any disturbance to the surface will remain stationary. In subcritical flow the flow is controlled from a downstream point and information is transmitted upstream. This condition leads to backwater effects. Supercritical flow is controlled upstream and disturbances are transmitted downstream.

Wave propagation can be used to illustrate these flow states: A stick placed in the water will create a V pattern of waves downstream. If flow is subcritical waves will appear in front of the stick. If flow is at critical waves will have a 45 o angle. If flow is supercritical no upstream waves will appear and the wave angle will be less than 45 o .

Note: Critical flow is unstable and often sets up standing waves between super and subcritical flow. When the actual water depth is below critical depth it is called supercritical because it is in a higher energy state. Likewise actual depth above critical depth is called subcritical because it is in a lower energy state.

Your Physicist

I will answer anything from the world of physics.

Froude number

What is froude number.

The Froude number is a dimensionless parameter used to determine the similarity between the flow of fluids at different scales. It is named after William Froude, a British engineer who studied the movement of ships and waves in water. The Froude number is defined as the ratio of the inertia forces to the gravitational forces acting on a fluid flow system. In simpler terms, it is a measure of how fast the fluid is moving relative to the depth of the fluid.

Calculation of Froude Number

The Froude number is calculated using the equation F = V/√(gL), where F is the Froude number, V is the velocity of the fluid, g is the acceleration due to gravity, and L is the characteristic length of the fluid flow. The characteristic length can be the depth of the fluid in a channel, the radius of a pipe, or the length of a ship. The Froude number is a dimensionless quantity, meaning that it has no units.

Applications of Froude Number

The Froude number is used in a variety of applications, including ship design, hydraulic engineering, and fluid mechanics. In ship design, the Froude number is used to determine the optimal speed of the ship and to ensure that the ship operates in a stable manner in different sea conditions. In hydraulic engineering, the Froude number is used to design channels and culverts for water flow. In fluid mechanics, the Froude number is used to study the behavior of fluids in different scenarios, such as the flow of water over a dam or the flow of air over an airplane wing.

Example of Froude Number in Action

One example of the use of the Froude number is in the design of a hydroelectric power plant. The Froude number is used to determine the optimal speed of the water flowing through the turbines. If the Froude number is too high, the water will not be able to flow smoothly through the turbines, resulting in inefficiency and damage to the equipment. If the Froude number is too low, the turbines will not be able to generate enough power. By calculating the Froude number, engineers can design the plant to operate at the most efficient speed and generate the maximum amount of power.

Search

www.springer.com The European Mathematical Society

  • StatProb Collection
  • Recent changes
  • Current events
  • Random page
  • Project talk
  • Request account
  • What links here
  • Related changes
  • Special pages
  • Printable version
  • Permanent link
  • Page information
  • View source

Froude number

One of the criteria for similarity of the motion of a liquid or gas, applicable in cases when the influence of gravitational forces is significant. The Froude number characterizes the relationship between the inertial and gravitational forces acting on an elementary volume of the liquid or gas. The Froude number is

$$\mathrm{Fr}=\frac{v^2}{g \ell},$$

where $v$ is the speed of the flow (or the speed of the moving body), $g$ is the gravitational acceleration and $\ell$ is the typical length of the flow or the body.

The Froude number was introduced by W. Froude (1870).

[a1] L.I. Sedov, "Similarity and dimensional methods in mechanics", Infosearch (1959) (Translated from Russian)
[a2] G. Birkhoff, "Hydrodynamics, a study in logic, fact and similitude", Princeton Univ. Press (1960)
  • This page was last edited on 4 January 2024, at 19:56.
  • Privacy policy
  • About Encyclopedia of Mathematics
  • Disclaimers
  • Impressum-Legal

Learning Materials

  • Business Studies
  • Combined Science
  • Computer Science
  • Engineering
  • English Literature
  • Environmental Science
  • Human Geography
  • Macroeconomics
  • Microeconomics
  • Froude Number

Explore the intricate concepts of Engineering Fluid Mechanics , starting with a comprehensive analysis of the Froude Number. This article provides in-depth explanations from understanding the theory, its importance, through to practical examples using the Froude Number Equation. It offers detailed insight into flow states, scrutinising both subcritical and critical flow. Additionally, readers can delve into advanced applications using a densimetric perspective and the relationship between the Froude Number and Dimensional Analysis . Unearth valuable knowledge and enhance your engineering proficiency with these real-world examples and case studies.

Millions of flashcards designed to help you ace your studies

  • Cell Biology

What is the significance of critical flow and when is a flow considered critical?

What does the Froude Number equation represent?

How are the subcritical and critical states of flow practically applied?

Can you describe the significance of the Froude Number in dimensional analysis?

What is the Froude Number in the context of fluid mechanics and when is the flow termed as subcritical?

What does the Froude Number represent in Engineering Fluid Mechanics?

How is the Froude Number applied in engineering fluid mechanics? Can you give an example?

What is the formula for calculating the Froude Number?

What are the main applications of the Froude Number in Engineering Fluid Mechanics?

How is the Froude Number used practically?

What is the Froude Number based on conceptually?

Review generated flashcards

to start learning or create your own AI flashcards

Start learning or create your own AI flashcards

  • Aerospace Engineering
  • Artificial Intelligence & Engineering
  • Automotive Engineering
  • Chemical Engineering
  • Design Engineering
  • Engineering Fluid Mechanics
  • Atmospheric Drag
  • Atmospheric Pressure
  • Atmospheric Waves
  • Axial Flow Pump
  • Bernoulli Equation
  • Boundary Layer
  • Boussinesq Approximation
  • Buckingham Pi Theorem
  • Capillarity
  • Cauchy Equation
  • Centrifugal Pump
  • Circulation in Fluid Dynamics
  • Colebrook Equation
  • Compressible Fluid
  • Continuity Equation
  • Continuous Matter
  • Control Volume
  • Convective Derivative
  • Coriolis Force
  • Couette Flow
  • Density Column
  • Dimensional Analysis
  • Dimensional Equation
  • Dimensionless Numbers in Fluid Mechanics
  • Dispersion Relation
  • Drag on a Sphere
  • Dynamic Pump
  • Dynamic Similarity
  • Dynamic Viscosity
  • Eddy Viscosity
  • Energy Equation Fluids
  • Equation of Continuity
  • Euler's Equation Fluid
  • Eulerian Description
  • Eulerian Fluid
  • Flow Over Body
  • Flow Regime
  • Flow Separation
  • Fluid Bearing
  • Fluid Density
  • Fluid Dynamic Drag
  • Fluid Dynamics
  • Fluid Fundamentals
  • Fluid Internal Energy
  • Fluid Kinematics
  • Fluid Mechanics Applications
  • Fluid Pressure in a Column
  • Fluid Pumps
  • Fluid Statics
  • Gas Molecular Structure
  • Gas Turbine
  • Hagen Poiseuille Equation
  • Heat Transfer Fluid
  • Hydraulic Press
  • Hydraulic Section
  • Hydrodynamic Stability
  • Hydrostatic Equation
  • Hydrostatic Force
  • Hydrostatic Force on Curved Surface
  • Hydrostatic Force on Plane Surface
  • Hydrostatics
  • Impulse Turbine
  • Incompressible Fluid
  • Internal Flow
  • Internal Waves
  • Inviscid Flow
  • Inviscid Fluid
  • Ion Thruster
  • Irrotational Flow
  • Jet Propulsion
  • Kinematic Viscosity
  • Kutta Joukowski Theorem
  • Lagrangian Description
  • Lagrangian Fluid
  • Laminar Flow in Pipe
  • Laminar vs Turbulent Flow
  • Laplace Pressure
  • Linear Momentum Equation
  • Liquid Molecular Structure
  • Mach Number
  • Magnetohydrodynamics
  • Mass Flow Rate
  • Material Derivative
  • Momentum Analysis of Flow Systems
  • Moody Chart
  • Navier Stokes Cartesian
  • Navier Stokes Cylindrical
  • Navier Stokes Equation in Spherical Coordinates
  • No Slip Condition
  • Non Newtonian Fluid
  • Nondimensionalization
  • Open Channel Flow
  • Orifice Flow
  • Pascal Principle
  • Plasma Parameters
  • Plasma Uses
  • Pneumatic Pistons
  • Poiseuille Flow
  • Positive Displacement Pump
  • Positive Displacement Turbine
  • Potential Flow
  • Prandtl Meyer Expansion
  • Pressure Change in a Pipe
  • Pressure Drag
  • Pressure Field
  • Pressure Head
  • Pressure Measurement
  • Pump Characteristics
  • Pump Performance Curve
  • Pumps in Series vs Parallel
  • Reaction Turbine
  • Relativistic Fluid Dynamics
  • Reynolds Experiment
  • Reynolds Number
  • Reynolds Transport Theorem
  • Rocket Propulsion
  • Rotating Frame of Reference
  • Rotational Flow
  • Sail Aerodynamics
  • Second Order Wave Equation
  • Shallow Water Waves
  • Shear Stress in Fluids
  • Shear Stress in a Pipe
  • Ship Propeller
  • Speed of Sound
  • Steady Flow
  • Steady Flow Energy Equation
  • Steam Turbine
  • Stokes Flow
  • Stream Function
  • Streamline Coordinates
  • Streamlines
  • Streamlining
  • Strouhal Number
  • Supersonic Flow
  • Surface Tension
  • Surface Waves
  • Torricelli's Law
  • Turbomachinery
  • Turbulent Flow in Pipes
  • Turbulent Shear Stress
  • Uniform Flow
  • Unsteady Bernoulli Equation
  • Unsteady Flow
  • Ursell Number
  • Varied Flow
  • Velocity Field
  • Velocity Potential
  • Velocity Profile
  • Velocity Profile For Turbulent Flow
  • Velocity Profile in a Pipe
  • Venturi Effect
  • Venturi Meter
  • Venturi Tube
  • Viscous Liquid
  • Volumetric Flow Rate
  • Wind Tunnel
  • Wind Turbine
  • Wing Aerodynamics
  • Womersley Number
  • Engineering Mathematics
  • Engineering Thermodynamics
  • Materials Engineering
  • Mechanical Engineering
  • Professional Engineering
  • Robotics Engineering
  • Solid Mechanics
  • What is Engineering

Understanding the Froude Number in Engineering Fluid Mechanics

In the field of Engineering Fluid Mechanics , the Froude Number is a significant non-dimensional parameter that plays a crucial role in analyzing various fluid flow scenarios. You'll come across this term quite often as you delve deeper into fluid mechanics, and understanding it is a stepping stone towards mastering this area of engineering.

The Froude Number, represented by \(Fr\), is a dimensionless number defining the ratio of the inertia force to the gravitational force acting on a fluid in motion. It is named after the British engineer and naval architect William Froude.

Definition of Froude Number

You might be curious to know more about the mathematics behind the definition of the Froude Number. For a body or surface moving in a fluid, or a flow past a stationary body or surface, the Froude Number is given by the formula:

  • \(v\) is the velocity of the object,
  • \(\sqrt{gL}\) is the speed of a gravity wave through the fluid,
  • \(g\) is the acceleration due to gravity,
  • and \(L\) is a characteristic length (for instance, the depth of flow or height of an object).

A Froude Number less than 1 indicates a subcritical flow condition where gravitational forces dominate. If the Froude Number is equal to 1, the flow is critical, and gravity and inertia forces are balanced. A Froude Number greater than 1 indicates a supercritical flow, dominated by inertia forces. This is quite important to understand as it helps in analyzing the types of flow in different situations.

Importance of Froude Number in Engineering Fluid Mechanics

The Froude Number carries a significant weight in Engineering Fluid Mechanics due to several reasons:

  • It plays a pivotal role in predicting the flow regime , whether the flow is laminar, transitional or turbulent.
  • In hydrodynamics, the Froude Number is used in the study of the stability of ships and boats, wave generation, and wave resistance.
  • The Froude Number is used for scaling fluid flow problems that involve a free surface. This implies that the Froude Number is highly valuable in designing flow models for rivers, channels, and hydraulic structures.
  • It’s also employed extensively in the field of environmental engineering while modelling wastewater and stormwater treatment systems.

Suppose there's a river with a flow velocity of 2 m/s and depth of 1.5 m. Considering the acceleration due to gravity as 9.81 m/s², the Froude Number would be \(\frac{2}{\sqrt{9.81 * 1.5}}\) = 0.65. The subcritical (or tranquil) flow state denotes that gravitational forces are the dominant factor in the flow regime of this river.

Comprehending the Froude Number Equation

Validation of the Froude Number and its incorporation into the broader subject of Engineering Fluid Mechanics is largely due to the mathematical equation it is represented by. This equation is pivotal in understanding the dynamics of fluid flow and its interaction with gravitational forces. The equation for the Froude Number is:

Here, \(Fr\) is the Froude Number, \(v\) is the velocity of the fluid, \(g\) is the acceleration due to gravity and \(L\) is a significant length associated with the problem, such as depth of flow.

Guidance on the Froude Number Derivation

The derivation of the Froude Number is based on the principle of dimensional analysis , which preferably involves Buckingham's Pi Theory. This method is widely used in fluid mechanics to generate dimensionless numbers.

To start with, consider the dimensions of velocity \([LT^{-1}]\), gravitational acceleration \([LT^{-2}]\) and length \([L]\). The aim here is to establish a dimensionless quantity by appropriately combining these variables.

Applying Buckingham's theorem with velocity, acceleration, and length as repeating variables, we find that there is one dimensionless group which could be formed. This leads to the combination \(v/\sqrt{gL}\), which is recognised as the Froude Number.

Indicative of its derivation, the Froude Number signifies the importance of inertial to gravitational forces in scaling fluid flow problems involving a free surface.

Applying the Froude Number Equation: Practical Examples

Being a versatile and effective tool, the Froude Number is used ubiquitously within fluid dynamics and engineering applications. Let's look at some practical scenarios and derive the Froude Number in each of them.

Example 1: Consider a water flow in a channel with a velocity of 3 m/s and a depth of 2 m. Given that the acceleration due to gravity is 9.81 m/s², the Froude Number can be calculated as follows:

In this case, as the Froude Number is less than 1, it denotes a subcritical flow condition, which signifies that gravitational forces dominate over inertial forces.

Example 2: Now consider a ship moving through the water with a velocity of 7.5 m/s. Let the significant length \(L\) in this case be the length of the ship's hull submerged in water, let's say 30 m. The Froude Number for this example would be:

This Froude Number much less than 1 indicates a highly subcritical flow, which implies that hydraulic jumps or sudden changes in the water elevation are less likely to occur around this ship.

In both these practical applications, and indeed in numerous others, the Froude Number plays a crucial role in helping us understand and predict the behaviour of fluid flows under the action of gravity and inertial forces.

Insight into Different Flow States: Subcritical and Critical Flow Froude Number

In the study of fluid mechanics, different flow states hold significant meaning. The Froude Number plays a critical role in defining these various states of flow including subcritical and critical flows. It helps in understanding whether the flow regime is dominated by inertial or gravitational forces.

Explanation of Froude Number Subcritical Flow

When addressing the states of flow, it’s essential to commence with the concept of subcritical flow. A subcritical flow happens when the Froude Number is less than one (\(Fr < 1\)). The name 'subcritical' signifies that the flow is under the critical point and is slow or tranquil. In this state, the flow is dominated by gravitational forces more than the inertial forces.

This feature of subcritical flow allows for small distortions or perturbations to propagate both downstream and upstream, which means that any change in the flow’s downstream conditions can influence the upstream flow. In subcritical flow, water profiles are often smooth and gentle showing no abrupt changes unless acted upon by an external force.

We can often see examples of subcritical flow in nature in slower-moving rivers or streams. Also, open channels that carry water at a steady and slow pace usually exhibit subcritical flow.

Consider a canal with a water velocity of 1 m/s and a flow depth of 1.5 m. Given the acceleration due to gravity as 9.81 m/s², the Froude Number for this canal would be calculated as \(Fr = \frac{1}{\sqrt{(9.81*1.5)}}\), which equals 0.26. This value of Froude Number less than 1 indicates a subcritical flow state, inferred as a slow, tranquil flow dominated by gravitational forces.

Scrutinising the Critical Flow Froude Number

Moving further along the spectrum of flow states, you reach critical flow , which occurs when the Froude Number equals one (\(Fr = 1\)). This is considered the dividing point between subcritical and supercritical flows. The critical flow is a state of balance where the inertial and gravitational forces acting on the fluid are equal.

Critical flow condition serves as the transition between the subcritical and supercritical states. Understandably, it is less common in natural or man-made flows given that any slight disturbance will cause the flow to shift into either the subcritical or supercritical state.

However, in practice, the critical flow is seen in certain fluid mechanics phenomena such as hydraulic jumps, where water abruptly transitions from supercritical to subcritical flow, or when analysing the maximum discharge capacity of a run-of-river hydropower plant or a spillway of a dam.

Consider a scenario where a waterfall has a velocity of 10 m/s just before it drops over a cliff edge which is 5 m high. Here, the Froude Number is given by \(Fr = \frac{10}{\sqrt{9.81*5}}\), which equals 1.41. As this value is greater than 1, the flow of the waterfall before the drop is in a supercritical state. However, at the very edge of the cliff where the waterfall begins to drop, the flow condition becomes critical before transitioning into a free fall, essentially a supercritical flow state.

In conclusion, the comprehension of Froude Number resulting in subcritical or critical flows is crucial in several fields of practical engineering, such as hydrology for designing channels, spillways and predicting flood levels, or naval architecture for designing ship hulls to minimise wave resistance.

Advanced Applications of Froude Number: The Densimetric Perspective

The traditional version of the Froude Number plays a significant role in various engineering applications where fluid flow and gravitational forces interact. However, when it comes to situations involving density differences within fluids, such as layered fluids or multiphase flows, a more advanced concept comes into play: the Densimetric Froude Number. This variant of the Froude Number significantly expands its usefulness, taking into account the density contrast in fluid flows, making it especially crucial in environmental and industrial applications.

Conceptualising the Densimetric Froude Number

The Densimetric Froude Number, often denoted as \(Fr_d\), pulls in the element of density difference between two fluids or between regions within a single fluid. The role of density variations becomes notable when we investigate stratified flows or multiphase flows, where lighter and heavier fluid layers or phases intermingle.

The Densimetric Froude Number is defined as: \(Fr_d = \frac{v}{\sqrt{g' \cdot L}}\).

In the above formula:

  • \(v\) is the characteristic velocity of the fluid,
  • \(L\) is a characteristic length,
  • and \(g'\) represents a reduced gravitational acceleration that introduces the density difference between the two layers or phases of the fluid and is given by \(g' = g \cdot \frac{\Delta \rho}{\rho_0}\), where \(\Delta \rho\) is the change in density and \(\rho_0\) is the reference density (often the density of the lighter fluid).

As with the original Froude Number, the Densimetric Froude Number is a dimensionless quantity. The interpretations linked to different values of \(Fr_d\) are similar to those for the traditional Froude Number. Yet, the inclusion of density contrast into the equation makes the Densimetric Froude Number significantly more relevant in scenarios of density stratified flows or multiphase flows.

It's fascinating to note that the Densimetric Froude Number has shown to be an essential parameter in the study of geophysical flows, particularly those linked to atmospheric and oceanographic phenomena. Here, density differences caused by temperature and salinity variations strongly impact the flow behaviour and dynamics, and the Densimetric Froude Number becomes a crucial tool for analysis and modelling.

Real-world Engineering Examples Utilising Densimetric Froude Number

The Densimetric Froude Number finds its applications in an array of real-world scenarios where fluid layers of different densities interact. Here are a couple of engineering cases in which it plays a key role:

Example 1: Atmospheric and Oceanic Flows: Probably the most widespread use of the Densimetric Froude Number is in geophysical fluid dynamics . Both atmospheric and oceanic flows often exhibit stratification due to temperature or salinity-induced density differences. By taking into account these density contrasts, the Densimetric Froude Number aids in appropriately scaling and investigating such phenomena. This can help in predicting weather patterns or ocean currents more accurately.

Example 2: Industrial Multiphase Flows: In industries, multiphase flows are quite common. Whether it's the oil and gas industry dealing with the simultaneous flow of oil, water, and gas in pipelines, or the food and chemical industries handling mixtures of liquids, solids, and gases, the Densimetric Froude Number becomes useful. It aids in characterising the flow regime and predicting phase distribution and pressure drop, thereby optimising the process performance.

Consider a pipeline in an oilfield carrying a mixture of crude oil (density = 850 kg/m³) and natural gas (density = 20 kg/m³). Let's say the mixture's velocity is 3 m/s and the pipeline diameter (characteristic length) is 0.1 m. Given the standard gravity as 9.81 m/s², we calculate the reduced gravity as \(g' = 9.81 \cdot \frac{(850 - 20)}{850}= 10.38\) m/s². The Densimetric Froude Number in this case can then be estimated as \(Fr_d = \frac{3}{\sqrt{10.38*0.1}}= 2.94\), which is greater than 1, indicating that the gas-oil flow in this pipeline is in a supercritical condition and dominated by inertia forces.

The value of the Densimetric Froude Number, in this situation and many like it, allows engineers to anticipate flow behaviour accurately and design effective operational strategies.

Froude Number and Dimensional Analysis

The study of fluid dynamics would be incomplete without the concept of dimensional analysis and the utilisation of dimensionless numbers, a chief one among them being the Froude Number. This section delves into the relationship between the Froude Number and dimensional analysis in the context of engineering fluid mechanics.

Relationship Between Froude Number and Dimensional Analysis

The practice of dimensional analysis is a powerful tool within physics and engineering disciplines, aiding not only in verifying equations and formulas but also in reducing complex physical phenomena to a simpler, more comprehensible form through dimensionless numbers. The Froude Number holds stature as one of these significant dimensionless numbers, primarily in studies involving gravity-driven fluid flows such as waves in oceans, rivers, and channels, where gravitational and inertia forces interact.

The Froude Number is defined as the ratio of inertial forces to gravitational forces: \(Fr = \frac{V}{\sqrt{gL}}\), where \(V\) is the characteristic velocity of the fluid, \(g\) is the acceleration due to gravity, and \(L\) represents a characteristic length.

This dimensionless number signifies the relative influence of these two forces on the flow behaviours. The fact that it's dimensionless makes it particularly useful when studying similar flow situations in differently scaled systems. For instance, water waves in a small laboratory tank or in an extensive ocean can be compared using the Froude Number, provided the flow is dynamically similar.

It is essential to realise that the construction of the Froude Number involves bringing together physical quantities of different dimensions (speed, length, gravity) using a square root operation. This is a classic example of how the process of dimensional analysis helps synthesise dimensionless quantities from dimensional ones.

Looking deeper into the subject, we find that fluid flow scenarios often involve more complexities beyond just inertia and gravity forces. For instance, viscosity and surface tension forces may become influential at smaller scales. Hence, in those situations, other dimensionless numbers like the Reynolds Number for inertial-viscous forces or the Weber Number for inertial-surface tension forces become significant. However, let it be noted that the Froude Number remains the go-to dimensionless number for large scale flows dominated by inertia and gravity forces.

Case Studies of Froude Number Dimensional Analysis in Engineering Fluid Mechanics

Understanding how the Froude Number and dimensional analysis work together can be best realised through practical case studies from the field of engineering fluid mechanics.

Case Study 1: Design of Ship Hulls: In naval architecture - the science of ship design - the hull shape plays a crucial role in a ship's resistance movement through water. The Froude Number is used as a significant parameter to ensure dynamic similarity between model tests in laboratories and real-world scenarios. For similar flows, if the ratios of inertial to gravitational forces (i.e., the Froude Numbers) of the model and the actual ship are equal, the wave patterns, wave resistances, and other hull performance characteristics will correspond. Therefore, making the accurate assessments using a small-scale model possible.

Case Study 2: River Modelling and Flood Prediction: Flood prediction and river management often rely on the construction of physical scale models of river segments. Here, the Froude Number enables the transfer of insights drawn from the scale models to the actual rivers. By ensuring that the Froude Number is the same in the model and reality, engineers can observe how changes in river flow characteristics like velocity, depth, and channel shape affect flood levels and consequently devise effective flood control measures.

For example, consider a situation where a large river is prone to flooding and engineers are designing a levee system to control it. Suppose they create a 1:100 scaled-down physical model of the river in a laboratory. In the model, if a particular flow velocity of 0.1 m/s results in safe water levels, they can use the Froude Number to determine the equivalent safe flow velocity in the actual river. If the model river depth (L) is 0.05 m, then the Froude Number in the model is \(Fr = \frac{0.1}{\sqrt{9.81*0.05}} = 0.45\). Assuming the same Froude Number in the actual river, with a depth of 5 m (100 times the model depth), the safe flow velocity can be calculated as \(V = Fr*\sqrt{9.81*5} = 0.45*\sqrt{9.81*5} = 1 m/s\). Hence, ensuring that the actual river flow velocity is maintained around this value will help achieve the desired safety against flooding, as indicated by the model study.

The above examples illustrate, not just the importance of the Froude Number and dimensional analysis in engineering, but also their real-world implications in mitigating risks and optimising system performance.

Froude Number - Key takeaways

  • The Froude Number is a dimensionless quantity which represents the ratio of the inertial force to gravitational force in free surface flow problems, calculated using the formula: \(Fr = \frac{v}{\sqrt{g \cdot L}}\).
  • Subcritical flow occurs when the Froude number is less than one (\(Fr < 1\)), suggesting that gravitational forces dominate over inertial forces. This flow state is typically slow and tranquil, with smooth water profiles and both upstream and downstream propagation of disturbances.
  • Critical flow is defined by a Froude number of one (\(Fr = 1\)), indicating a balance of inertial and gravitational forces on the fluid. It serves as the transition point between subcritical and supercritical flow states.
  • The Densimetric Froude Number accounts for density differences within fluids in scenarios like layered or multiphase flows. It is calculated using the formula: \(Fr_d = \frac{v}{\sqrt{g' \cdot L}}\) where \(g'\) represents a reduced gravitational acceleration considering density difference.
  • The Froude Number is derived through the process of dimensional analysis, specifically using Buckingham's Pi theorem. This helps establish the Froude Number as a significant dimensionless group, helping to scale and analyze fluid flow problems involving a free surface.

Flashcards in Froude Number 15

Critical flow is a balance state when the Froude Number equals one. This implies that inertial and gravitational forces acting on the fluid are equal. This condition is the transition point between subcritical and supercritical flows.

The Froude Number equation represents the ratio of inertial to gravitational forces in fluid flow problems involving a free surface. The variables are velocity of the fluid, acceleration due to gravity and significant length related to the problem.

The understanding of subcritical and critical flows is essential in practical engineering fields like hydrology for designing channels, spillways, flood level prediction or naval architecture for designing ship hulls to minimise wave resistance.

The Froude Number is a powerful tool in dimensional analysis as it helps to represent the interaction of gravitational and inertia forces in fluid flows. Its dimensionless nature allows for comparing similar flow situations in differently scaled systems. It's particularly useful in studying large-scale flows dominated by inertia and gravity.

The Froude Number in fluid mechanics defines different states of flow such as subcritical and critical. A flow is considered subcritical when the Froude Number is less than one, meaning the flow regime is dominated by gravitational forces more than the inertial forces.

The Froude Number, symbolized by \(Fr\), is a non-dimensional parameter in Engineering Fluid Mechanics that defines the ratio of the inertia force to the gravitational force acting on a fluid in motion.

Froude Number

Learn with 15 Froude Number flashcards in the free StudySmarter app

We have 14,000 flashcards about Dynamic Landscapes.

Already have an account? Log in

Frequently Asked Questions about Froude Number

Test your knowledge with multiple choice flashcards.

Froude Number

Join the StudySmarter App and learn efficiently with millions of flashcards and more!

Keep learning, you are doing great.

Discover learning materials with the free StudySmarter app

1

About StudySmarter

StudySmarter is a globally recognized educational technology company, offering a holistic learning platform designed for students of all ages and educational levels. Our platform provides learning support for a wide range of subjects, including STEM, Social Sciences, and Languages and also helps students to successfully master various tests and exams worldwide, such as GCSE, A Level, SAT, ACT, Abitur, and more. We offer an extensive library of learning materials, including interactive flashcards, comprehensive textbook solutions, and detailed explanations. The cutting-edge technology and tools we provide help students create their own learning materials. StudySmarter’s content is not only expert-verified but also regularly updated to ensure accuracy and relevance.

Froude Number

StudySmarter Editorial Team

Team Engineering Teachers

  • 17 minutes reading time
  • Checked by StudySmarter Editorial Team

Study anywhere. Anytime.Across all devices.

Create a free account to save this explanation..

Save explanations to your personalised space and access them anytime, anywhere!

By signing up, you agree to the Terms and Conditions and the Privacy Policy of StudySmarter.

Sign up to highlight and take notes. It’s 100% free.

Join over 22 million students in learning with our StudySmarter App

The first learning app that truly has everything you need to ace your exams in one place

  • Flashcards & Quizzes
  • AI Study Assistant
  • Study Planner
  • Smart Note-Taking

Join over 22 million students in learning with our StudySmarter App

Fluid Mechanics
Hydraulic Jump through a Sluice Gate
Introduction

A hydraulic jump is a sudden dissipation of energy caused by a change from super-critical to sub-critical flow. The concept is very similar to sudden expansion in pipe flow, except that hydraulic jumps occur in open-channel flow. At the start of the jump, the flow height will begin to increase, and the velocity will slow creating an area of turbulence. At the end of the jump, the flow height will level off again, and the fluid will continue flowing smoothly.

Before further discussion of hydraulic jumps, it is necessary to define sub-critical and super-critical flow. Fluid flowing in an open channel must have some minimum amount of energy, E min . E min = 1.5 y c , where y c is defined as the critical depth. The stream velocity at y c is defined as V c or critical velocity. A large flow height with V < V c is called sub-critical flow. In sub-critical flow, waves can travel upstream since the wave speed is greater than the free stream velocity. Super-critical flow is defined as a small small flow height causing V > V c (by conservation of mass). Waves always travel downstream in super-critical flow as the free stream velocity is greater than the wave speed.

Hydraulic jumps dissipate a large amount of energy in open channel flows. This makes hydraulic jumps very useful in dam and spillway designs. Many times assistance is needed to make jumps occur at desired locations near spillways. Increasing surface roughness, adding a baffle wall, or sloping the basin floor can all help force a hydraulic jump. The major factor behind a hydraulic jump is the Froude Number, Fr. The best design range for the Froude number is 4.5 to 9.0. In this range, a well-balanced steady jump will occur with a large amount of energy dissipation. A Froude number of 2.5 to 4.5 is the worst design range, as the jump in this range will create large waves that could cause structural damage.

The performance of a hydraulic jump depends mainly on the value of the Froude number. The Froude number is defined as

V1
(gy1)1/2

where V 1 is the stream velocity before the jump, y 1 is the flow height before the jump, and g is the universal gravitational constant.

Fr Number Jump Description
< 1.0 No jump since flow is already sub-critical
1.0 to 1.7 An undular jump, with about 5% energy dissipation
1.7 to 2.5 A weak jump with 5% to 15% energy dissipation
2.5 to 4.5 Unstable, oscillating jump, with 15% to 45% energy dissipation
4.5 to 9.0 Stable, steady jump with 45% to 70% energy dissipation
> 9.0 Rough, strong jump with 70% to 85% energy dissipation

The two parameters that can easily be varied to change the Froude number are the stream velocity and the flow height. In this experiment, the velocity is varied by varying the speed of the pump. The height of the flow is varied usingf a sluice gate (a baffle with some space underneath that allows the flow to travel through at the set height).

  • The bump was constructed from 6" section of 3" diameter PVC pipe. The pipe was cut in half producing a three dimensional semi - circle. Two holes were drilled through the pipe - section. Two threaded metal rods were put through the holes and fastened with nuts. The metal rods were bent at an angle of 90o and attached to the sluice gate apparatus. Holes were drilled into the Plexiglas cover plate at a distance of 5 cm increments such that the bump will be directly under the sluice gate initially.
  • The Plexiglas cover plate containing the sluice gate was inserted into position on the top of the water channel with an initial gate height of 2.5 cm from the bottom of the water channel.
  • The water channel was filled half way with water.
  • The Toshiba Frequency Controller was turned on.
  • The speed of the pump was increased using the Toshiba Frequency Controller until a hydraulic jump was formed.
  • The height of the water before and after the gate was measured.
  • The distance from the gate to the hydraulic jump was measured.
  • The laser velocimeter was used to calculate velocity before the hydraulic jump.
  • A photograph was taken of the hydraulic jump using the digital camera.
  • The pump was turned off.
  • The bump was inserted into the water channel at various distances in accordance to the holes drilled in the cover plate. The steps 4 to 10 were repeated at each distance from the sluice gate.
  • Steps 4 to 11 were repeated at several sluice gate heights (increments of 1 cm) and increased pump speeds.
Download Files
FileTitleTypeSize
Hydraulic Jump Data Microsoft Excel 14K

Chanson, H. Hydraulic Design of Stepped Cascades, Channels, Weirs, and Spillways . Pergamon, 1994.

Vischer, Daniel L. and Hager, Willi H. ed. Energy Dissipators . A. A. Balkema / Rotterdam / Brookfield, 1995.

White, Frank M. Fluid Mechanics 4th ed . McGraw-Hill, 1999

  • Calculate the Froude number for each configuration.
  • Plot the Froude number and bump distance at each velocity
  • Report the results and discuss.

Talk to our experts

1800-120-456-456

  • Froude Number

ffImage

What is Froude Number?

In the dimension analysis, the dimensionless quantity is the quantity that does not have physical dimensions. These dimensionless quantities are defined as the product or ratio of the quantities with dimensions. For example, strain is the measure of deformation, in other words, it can be written as the ratio of change in length over the initial length. Since the unit is ‘L’ in both cases thus a strain is dimensionless. 

The dimensionless numbers are used to analyze the prototype models. These numbers are Reynolds numbers, Weber’s number, and Froude’s number. In situations where the gravitational force is important, the Froude number governs dynamic similarity. Some of the examples include, flow through the open channels, flow over the spillway of the dam, etc. 

Froude Number Formula

Froude number is represented by Fr, in fluid mechanics, and in hydrology, Fr is a dimensionless quantity that is used in the indication of the influence of gravity on the motion of fluids. It is expressed as,

Froude No = \[\frac{v}{\sqrt{gd}}\]

Here, ‘d’ represents the depth of flow, 

‘g’ is the acceleration due to gravity,

‘v’ is the celerity or gravity of the small surface wave . 

When Froude No. is at a critical point that is Fr = 1, then the velocity of the flow is equal to the velocity of surface waves, which is also known as the Critical flow Froude number. 

If the Fr is less than one, then this means that the small surface waves can move in an upstream direction or it is considered as fluvial motion. 

If Fr is greater than one, then this means that the small surface waves can move in the downstream direction or it is considered as torrential flow motion. The Froude number enters into the hydraulic jump that occurs in certain conditions and along with the Reynolds number. 

Froude Definition

Froude’s number is named after the scientist William Froude. It depends on the speed-length ratio. It has some analogy with the Mach number. In the case of theoretical fluid dynamics, it is not considered frequently because usually the equations are considered in the high Froude limit in the case of the negligible external field. 

In naval architecture, the Froude number is the significant figure that is used in the determination of the resistance of the partially submerged object that is moving through the water. In the open channel flows, the ratio of the flow velocity to the square root of the product of acceleration due to gravity to the depth of the flow. 

William Froude used a series of scale models to measure the resistance of each module when it is towed at the given speed. This concept was first introduced by Frederic Reech and is used for testing ships and propellers. But the speed-length ratio was originally derived by Fraude in his Law of Comparison. Where the dimensional terms consist of 

Speed - length ratio = \[\frac{u}{\sqrt{LWL}}\]

Where u is the flow speed and LWL indicates the length of the waterline. It is converted into a non-dimensional term. It is also known as the Reech–Froude number in the region of France.

Froude Number Applications

1. Cauchy Momentum Equation: The characteristic length r0 and characteristic velocity u0 are to be defined to make the equations dimensionless. The dimensionless variables should be of one order. 

ρ* ≡ \[\frac{\rho }{\rho _0}\], u* ≡ \[\frac{u}{u_0}\], r* ≡ \[\frac{r}{r_0}\], t* ≡ \[\frac{u_0}{r_0}\]t, g* ≡ \[\frac{g}{g_0}\], σ* ≡ \[\frac{\sigma }{\sigma _0}\] 

If we substitute these inverse relations in the Euler momentum equations we get,

Froude No = \[\frac{u_0}{\sqrt{g_0r_0}}\]

Thus Cauchy momentum equation is,

\[\frac{Du}{Dt}\] + Eu\[\frac{1}{\rho }\] ▽ . σ = \[\frac{1}{Fr^2}\]g 

In the high Froude limit, the Cauchy type of equations is known as free equations, where Fr → ∞ and in the lower Euler limit, where Eu → 0, the Cauchy momentum equation becomes in the homogeneous equation called the Burgers equation.

\[\frac{\partial u}{\partial t}\] + ▽ . (\[\frac{1}{2}\]u ⊗ u) = \[\frac{1}{Fr^2}\]g

2. Euler Momentum Equation: It is a Cauchy momentum equation along with the Pascal law with the relation of stress constitutive.

3. Incompressible Navier–Stokes Momentum Equation: It is a Cauchy momentum equation along with the Pascal law and Stoke’s law with the relation of stress constitutive.

4. Ship Hydrodynamics: In the applications of marine hydrodynamics, Froude number is usually referred to as Fn, 

Fn L = \[\frac{u}{\sqrt{gL}}\]

Here, u is considered as the relative flow velocity in between the sea and ship, g is the acceleration due to gravity and L is the length of the ship at the level of the waterline, it is also denoted as Lwl. It is the important parameter that is calculated for the resistance or ship drag in respect to the wave-making resistance.

In the planning crafts, while the waterline length is too speed-dependent, the Froude number can be defined as displacement Froude number, and thus the reference length is taken from the volumetric displacement of the hull, it is represented as:

Fn V = \[\frac{u}{\sqrt{g^3}\sqrt{v}}\]

5. Shallow Water Waves: The shallow waterways such as hydraulic jumps and tsunamis, U is considered as the average flow velocity, where this U is averaged to the cross-section that is perpendicular to the flow direction. Thus the wave velocity c is defined as:

c = \[\sqrt{g\frac{A}{B}}\], hence the Froude number in the shallow water will be, 

Fr = \[\frac{u}{\sqrt{g\frac{A}{B}}}\]

For the uniform depth and rectangular cross-section, the Froude number is given as,

Fr = \[\frac{u}{\sqrt{gd}}\]

If Fr ≈ 1 then the flow is critical, 

If the Fr < 1 then the flow is Subcritical, 

If the Fr > 1 then the flow is Supercritical. 

6. Wind Engineering: In dynamically sensitive structures, considering the wind effects is necessary for the simulation of the combined effect of the vibrating mass along with the fluctuating force of the wind. When simulating the hot smoke plumes when combined with the natural wind, the scaling of the Froude number is necessary for the maintenance of the correct balance between the buoyancy forces and the momentum of the wind. 

Hydraulic Jump

In the science of hydraulics, the hydraulic jump is the phenomenon that is frequently observed in the case of open channel flows that include rivers and spillways. When the liquid is at a high velocity and if it discharges to the zone of lower velocity then the rise occurs in the region of the liquid surface. This rapidly flowing water abruptly slows down and increases its height, thus converting some of the flow from kinetic energy to potential energy during this conversion some of the energy is lost in the form of heat of turbulence. 

This phenomenon is dependent on the initial speed of the fluid, if the initial speed is said to be lower than that of the critical speed then there is no jump observed in the liquid, thus in such cases, the transition appears as an undulating wave. The hydraulic jump is seen in both dynamic form and stationary form. The stationary form is known as a hydraulic jump, and in the dynamic form, it is known as positive surge or hydraulic jump of translation. 

The final equation in the representation of the hydraulic jump consists of Froude No, it is also known as Belanger equation, it is represented as,

\[\frac{h_1}{h_2}=\frac{\sqrt{1+8Fr^2}-1}{2}\]

If the Fr value is greater than one, then the initial flow represents the supercritical flow, and if the Fr is less than one, then it represents the subcritical flow. 

The dam should be designed in such a way that the energy of the fast-flowing stream of the spillway should be dissipated partially to avoid the erosion of the streambed downstream of the spillway at the end this leads to failure of the dam. An apron is designed to withstand the hydraulic forces and prevent the local cavitation to avoid erosion. Thus to design these the engineers select a point where the hydraulic jump will occur. To trigger the hydraulic jump, the apron is designed in such a way that the flat slope of this retards the flow of the water from the spillway. There are two methods to induce the jump:

If the downstream channel restricts the downstream flow, then the water backs up onto the foot of the spillway, thus the downstream water level is used to determine the location of the hydraulic jump.

At some distance, the spillway continues to drop the flow, but this flow will be changed to prevent the support of the supercritical flow since the depth of the lower subcritical flow region is required to identify the location of the jump. 

In both of these cases, the final depth of the water can be determined by downstream characteristics, 

\[h_0=\frac{h_1}{2}(-1+\sqrt{1+8Fr_2^2}\]

Where Fr represents the upstream Froude number, g is the acceleration due to gravity, h is the height of the fluid.

Based on the Froude number hydraulic jump is classified as:

If the Fr is in between 1.0 to 1.7, then it is known as undular jump.

If the Fr is between 1.7 to 2.5, then it is known as a weak jump.

If the Fr is in between 2.5 to 4.5 it is known as an oscillating jump.

If the Fr is in between 4.5 to 9.0 then it is known as a steady jump.

Froude's model law :

A model's Froude number should be the same as a prototype's Froude number. It is used in situations where gravity is simply acting as an overriding force to govern flow as well as inertia force.

Uses of Froude Number :

Following are some of the uses of Froude Number:

It is significant in fluid dynamics problems when the fluid's weight is a substantial force.

It is taken into account while examining flow in spillways, channel flows, weirs, rivers, and ship design.

It is useful for calculating hydraulic jumps and building hydraulic structures when gravity and inertial forces are in control.

Bedforms' characteristics shall be determined

It is used to overcome the resistance of an object moving through water.

Fr. compares the resistance caused by a wave among bodies of various shapes and sizes.

It is useful in low-water waves, wind engineering, and ship hydrodynamics.

In between the bodies of various sizes and shapes, the Froude number is used to determine wave-making resistance. In the case of free surface flow, the nature of the flow is determined by the value of the Froude number that is greater than or less than the unity. The line of critical flow is usually in the kitchen or bathroom sinks. 

When we keep it unplugged and allow the faucet to run, the place where the stream hits the sink, that flow is considered as the supercritical flow. The stream hugs the surface and moves very quickly. The outer edge of the flow pattern is subcritical, the boundary between these two areas is known as the hydraulic jump. This jump starts only when the Froude number is 1.

arrow-right

FAQs on Froude Number

1. After whom is Froude’s number named?

The number is named after the scientist William Froude. William Froude measured the resistance of each module while it was towed at the specified speed using a series of scale models. This concept was developed by Frederic Reech and is used to test ships and propellers. The speed-length ratio, on the other hand, was discovered by Fraude in his Law of Comparison. It is determined by the speed-length ratio. It bears some resemblance to the Mach number. It is not typically studied in theoretical fluid dynamics since the equations are normally considered at the high Froude limit in the event of a negligible external field.

2. What is a hydraulic jump?

The hydraulic jump is a behavior that is commonly observed in open channel flows such as rivers and spillways in the study of hydraulics. The rise happens at the liquid surface when the liquid is at a high velocity and discharges to a zone of lower velocity. This fast-moving water suddenly pauses and rises, transforming some of the flow from kinetic to potential energy. Some of the energy is wasted in the form of heat of turbulence during this conversion.

3. What are the manifestations of hydraulic jump?

Hydraulic jumps have two basic manifestations, and formerly, separate terminology has been used for each. The mechanisms underlying them, however, are identical since they are merely variations of each other observed from various frames of reference, and so the physics and analysis tools can be applied to both types.

The many manifestations are as follows:

The stationary hydraulic jump — swiftly flowing water transitions to slowly moving water in a stationary jump.

The tidal bore - a wall or undulating wave of water pushes upstream against water flowing downstream. When a frame of reference moves with the wavefront, the wavefront becomes stationary relative to the frame and shows the same key behaviour as the stationary leap.

4. Mention the Types of Jump Depending on the Froude Number.

Fr is in between 1.0 to 1.7, then it is an undular jump.

Fr is between 1.7 to 2.5, then it is a weak jump.

Fr is in between 2.5 to 4.5 it is an oscillating jump.

Fr is in between 4.5 to 9.0 then it is a steady jump.

Information

  • Author Services

Initiatives

You are accessing a machine-readable page. In order to be human-readable, please install an RSS reader.

All articles published by MDPI are made immediately available worldwide under an open access license. No special permission is required to reuse all or part of the article published by MDPI, including figures and tables. For articles published under an open access Creative Common CC BY license, any part of the article may be reused without permission provided that the original article is clearly cited. For more information, please refer to https://www.mdpi.com/openaccess .

Feature papers represent the most advanced research with significant potential for high impact in the field. A Feature Paper should be a substantial original Article that involves several techniques or approaches, provides an outlook for future research directions and describes possible research applications.

Feature papers are submitted upon individual invitation or recommendation by the scientific editors and must receive positive feedback from the reviewers.

Editor’s Choice articles are based on recommendations by the scientific editors of MDPI journals from around the world. Editors select a small number of articles recently published in the journal that they believe will be particularly interesting to readers, or important in the respective research area. The aim is to provide a snapshot of some of the most exciting work published in the various research areas of the journal.

Original Submission Date Received: .

  • Active Journals
  • Find a Journal
  • Proceedings Series
  • For Authors
  • For Reviewers
  • For Editors
  • For Librarians
  • For Publishers
  • For Societies
  • For Conference Organizers
  • Open Access Policy
  • Institutional Open Access Program
  • Special Issues Guidelines
  • Editorial Process
  • Research and Publication Ethics
  • Article Processing Charges
  • Testimonials
  • Preprints.org
  • SciProfiles
  • Encyclopedia

water-logo

Article Menu

froude experiment

  • Subscribe SciFeed
  • Recommended Articles
  • Google Scholar
  • on Google Scholar
  • Table of Contents

Find support for a specific problem in the support section of our website.

Please let us know what you think of our products and services.

Visit our dedicated information section to learn more about MDPI.

JSmol Viewer

Numerical analysis of flow characteristics and energy dissipation on flat and pooled stepped spillways.

froude experiment

1. Introduction

2. material and methods, 2.1. physical model, 2.2. numerical simulation, 2.2.1. turbulence model, 2.2.2. realizable k-ε model, 2.3. computational setup, 2.3.1. boundary conditions, 2.3.2. gird and mesh assessment, 3. results and discussion, 3.1. validation of numerical model, 3.2. flow behavior and water surface profiles, 3.3. water surface profile for geometry case 1, 3.4. water surface profile for geometry case 2, 3.5. water surface profile for geometry case 3, 3.6. water surface profile for geometry case 4, 3.7. velocity and pressure distribution, 4. conclusions, author contributions, data availability statement, acknowledgments, conflicts of interest.

  • Ghaderi, A.; Abbasi, S.; di Francesco, S. Numerical Study on the Hydraulic Properties of Flow over Different Pooled Stepped Spillways. Water 2021 , 13 , 710. [ Google Scholar ] [ CrossRef ]
  • Ashoor, A.; Riazi, A. Stepped Spillways and Energy Dissipation: A Non-Uniform Step Length Approach. Appl. Sci. 2019 , 9 , 5071. [ Google Scholar ] [ CrossRef ]
  • Limantara, L.M.; Malindo, D.; Juwono, P.T.; Hendrawan, A.P. Modelling Spillway for Flood Control Optimization in Embankment Dry Dam. J. Law Sustain. Dev. 2024 , 12 , e3277. [ Google Scholar ] [ CrossRef ]
  • Matos, J.; Meireles, I. Hydraulics of Stepped Weirs and Dam Spillways: Engineering Challenges, Labyrinths of Research. In Proceedings of the Hydraulic Structures and Society—Engineering Challenges and Extremes, Brisbane, Australia, 25–27 June 2014; pp. 1–30. [ Google Scholar ]
  • James, C.S.; Main, A.G.; Moon, J. Enhanced Energy Dissipation in Stepped Chutes. Proc. Inst. Civ. Eng.-Water Marit. Eng. 2001 , 148 , 277–280. [ Google Scholar ] [ CrossRef ]
  • Felder, S.; Chanson, H. Simple Design Criterion for Residual Energy on Embankment Dam Stepped Spillways. J. Hydraul. Eng. 2016 , 142 , 04015062. [ Google Scholar ] [ CrossRef ]
  • Mason, P.J.; Hinks, J.L. Security of Stepped Masonry Spillways: Lessons from Ulley Dam. Dams Reserv. 2008 , 18 , 5–8. [ Google Scholar ] [ CrossRef ]
  • Souli, H.; Ahattab, J.; Bensallam, S. Influence of a Modified Weir Profile on Velocity Field and Dissipation Rate in Stepped Spillways: A Comparative Study Using Physical Models and Computational Fluid Dynamics. JAFM 2024 , 17 , 1874–1884. [ Google Scholar ] [ CrossRef ]
  • Farooq, U.; Taha Bakheit Taha, A.; Tian, F.; Yuan, X.; Ajmal, M.; Ullah, I.; Ahmad, M. Flood Modelling and Risk Analysis of Cinan Feizuo Flood Protection Area, Huaihe River Basin. Atmosphere 2023 , 14 , 678. [ Google Scholar ] [ CrossRef ]
  • Ghaderi, A.; Abbasi, S. Experimental and Numerical Study of the Effects of Geometric Appendance Elements on Energy Dissipation over Stepped Spillway. Water 2021 , 13 , 957. [ Google Scholar ] [ CrossRef ]
  • Salmasi, F.; Abraham, J. Genetic Algorithms for Optimizing Stepped Spillways to Maximize Energy Dissipation. Water Supply 2022 , 22 , 1255–1274. [ Google Scholar ] [ CrossRef ]
  • Felder, S.; Chanson, H. Energy Dissipation down a Stepped Spillway with Nonuniform Step Heights. J. Hydraul. Eng. 2011 , 137 , 1543–1548. [ Google Scholar ] [ CrossRef ]
  • Bombardelli, F.A.; Meireles, I.; Matos, J. Laboratory Measurements and Multi-Block Numerical Simulations of the Mean Flow and Turbulence in the Non-Aerated Skimming Flow Region of Steep Stepped Spillways. Environ. Fluid Mech. 2011 , 11 , 263–288. [ Google Scholar ] [ CrossRef ]
  • Alqarni, M.S.; Memon, A.A.; Anwaar, H.; Usman; Muhammad, T. The Forced Convection Analysis of Water Alumina Nanofluid Flow through a 3D Annulus with Rotating Cylinders via Κ−ε Turbulence Model. Energies 2022 , 15 , 6730. [ Google Scholar ] [ CrossRef ]
  • Nikmehr, S.; Aminpour, Y. Numerical Simulation of Hydraulic Jump over Rough Beds. Period. Polytech. Civ. Eng. 2020 , 64 , 396–407. [ Google Scholar ] [ CrossRef ]
  • Zabaleta, F.; Bombardelli, F.A.; Márquez Damián, S. Numerical Modeling of Self-Aerated Flows: Turbulence Modeling and the Onset of Air Entrainment. Phys. Fluids 2024 , 36 , 043337. [ Google Scholar ] [ CrossRef ]
  • Ghaderi, A.; Dasineh, M.; Aristodemo, F.; Aricò, C. Numerical Simulations of the Flow Field of a Submerged Hydraulic Jump over Triangular Macroroughnesses. Water 2021 , 13 , 674. [ Google Scholar ] [ CrossRef ]
  • Ma, X.; Zhang, J.; Hu, Y. Numerical Investigation of the Flow Pattern and Energy Dissipation of Interval-Pooled Stepped Spillways. In Proceedings of the E-Proceedings of the 38th IAHR World Congress, Panama City, Panama, 1–6 September 2019; pp. 2521–2530. [ Google Scholar ]
  • Morovati, K.; Eghbalzadeh, A.; Soori, S. Study of Energy Dissipation of Pooled Stepped Spillways. Civ. Eng. J. 2016 , 2 , 208–220. [ Google Scholar ] [ CrossRef ]
  • Li, J.; Zhou, Y.; Shu, Z.; Mao, X.; Xu, M.; Ye, Z.; Ye, J. A Flow-3d-Based Numerical Simulation of the Energy Dissipation Characteristics of Pooled Stepped Spillways. In Proceedings of the 2022 8th IEEE International Conference on Hydraulic and Civil Engineering: Deep Space Intelligent Development and Utilization Forum (ICHCE), Xi’an, China, 25–27 November 2022; pp. 989–994. [ Google Scholar ]
  • Li, S.; Yang, J.; Li, Q. Numerical Modelling of Air-Water Flows over a Stepped Spillway with Chamfers and Cavity Blockages. KSCE J. Civ. Eng. 2020 , 24 , 99–109. [ Google Scholar ] [ CrossRef ]
  • Saleh, S.M.; Muhammad, S.H.; Abo, A.A. Effect of Pooled and Flat Stepped Spillway on Energy Dissipation Using Computational Fluid Dynamics. Tikrit. J. Eng. Sci. 2022 , 29 , 75–79. [ Google Scholar ] [ CrossRef ]
  • Hantoosh, S.H.; Shamkhi, M.S. Discharge Coefficient and Energy Dissipation on Stepped Weir. Open Eng. 2023 , 13 , 20220427. [ Google Scholar ] [ CrossRef ]
  • Pandey, B.R.; Megh Raj, K.C.; Crookston, B.; Zenz, G. Numerical Investigation of Different Stepped Spillway Geometries over a Mild Slope for Safe Operation Using Multiphase Model. Water 2024 , 16 , 1635. [ Google Scholar ] [ CrossRef ]
  • Guenther, P.; Felder, S.; Chanson, H. Flow Aeration, Cavity Processes and Energy Dissipation on Flat and Pooled Stepped Spillways for Embankments. Environ. Fluid Mech. 2013 , 13 , 503–525. [ Google Scholar ] [ CrossRef ]
  • Salmasi, F.; Abraham, J. Effect of Slope on Energy Dissipation for Flow over a Stepped Spillway. Water Supply 2022 , 22 , 5056–5069. [ Google Scholar ] [ CrossRef ]
  • Peng, Y.; Zhang, X.; Yuan, H.; Li, X.; Xie, C.; Yang, S.; Bai, Z. Energy Dissipation in Stepped Spillways with Different Horizontal Face Angles. Energies 2019 , 12 , 4469. [ Google Scholar ] [ CrossRef ]
  • Gubashi, K.R.; Mulahasan, S.; Hacheem, Z.A.; Rdhaiwi, A.Q. Effect of the Stepped Spillway Geometry on the Flow Energy Dissipation. Civ. Eng. J. 2024 , 10 , 145–158. [ Google Scholar ] [ CrossRef ]
  • Ma, X.; Zhang, J.; Hu, Y. Analysis of Energy Dissipation of Interval-Pooled Stepped Spillways. Entropy 2022 , 24 , 85. [ Google Scholar ] [ CrossRef ]
  • Hunt, S.L.; Kadavy, K.C. The Effect of Step Height on Energy Dissipation in Stepped Spillways. In Proceedings of the World Environmental and Water Resources Congress 2009, Kansas City, MO, USA, 17–21 May 2009; American Society of Civil Engineers: Kansas City, MO, USA, 2009; pp. 1–11. [ Google Scholar ]
  • Daneshfaraz, R.; Sadeghi, H.; Ghaderi, A.; Abraham, J.P. Characteristics of Hydraulic Jump and Energy Dissipation in the Downstream of Stepped Spillways with Rough Steps. Flow Meas. Instrum. 2024 , 96 , 102506. [ Google Scholar ] [ CrossRef ]
  • Hirt, C.W.; Nichols, B.D. Volume of Fluid (VOF) Method for the Dynamics of Free Boundaries. J. Comput. Phys. 1981 , 39 , 201–225. [ Google Scholar ] [ CrossRef ]
  • Kaouachi, A.; Carvalho, R.F.; Lopes, P.; Benmamar, S.; Gafsi, M. Numerical Investigation of Alternating Skimming Flow over a Stepped Spillway. Water Supply 2021 , 21 , 3837–3859. [ Google Scholar ] [ CrossRef ]
  • Kositgittiwong, D.; Chinnarasri, C.; Julien, P.Y. Numerical Simulation of Flow Velocity Profiles along a Stepped Spillway. Proc. Inst. Mech. Eng. Part E J. Process Mech. Eng. 2013 , 227 , 327–335. [ Google Scholar ] [ CrossRef ]
  • Khuntia, J.R.; Devi, K.; Khatua, K.K. Prediction of Depth-Averaged Velocity in an Open Channel Flow. Appl. Water Sci. 2018 , 8 , 172. [ Google Scholar ] [ CrossRef ]
  • Manogaran, T.; Mohd Arif Zainol, M.R.R.; Wahab, M.K.A.; Abdul Aziz, M.S.; Zahari, N.M. Assessment of Flow Characteristics along the Hydraulic Physical Model of a Dam Spillway. J. Civ. Eng. Sci. Technol. 2022 , 13 , 69–79. [ Google Scholar ] [ CrossRef ]

Click here to enlarge figure

ParameterSetting
Solution methodPressure–velocity coupling
Discretization schemeSIMPLE algorithm
MomentumSecond order upwind
PressureModified body force-weighted
Volume fractionCompressive
Turbulent kinetic energySecond order upwind
Turbulent dissipation rateSecond order upwind
Transient formulationFirst order implicit
Viscous modelk-ε model standard
Near wall treatmentStandard wall function
Time step size0.0001 s
Residual for all equation parameters0.0001
Quality MetricDescriptionAcceptableGrid Size
SkewnessMeasures the deviation of cell shapes from the ideal≤0.850.00624
Orthogonal QualityEvaluates the orthogonality of cell faces and edges≥0.10.00624
Aspect RatioRatio of the longest edge to the shortest edge of a cell≤3 for most applications0.00624
(l/s)y (m) y (m)y /hy /hh (m)h (m)l (m)y /y y /y L L L L Re
450.0530.0540.880.90 0.060.060.1212.5011.765.655.023.433.233.5 × 10
500.0570.0580.950.960.060.060.1212.7912.2505.865.293.663.453.9 × 10
540.0620.0601.031.000.060.060.1212.9012.2825.935.633.923.634.2 × 10
V (m/s)uy (m)y /hy /y Fr Fr
0.451.300.0520.8663.4660.6302.322
0.501.500.0560.9335.1870.6742.570
0.551.730.0580.9666.0000.7293.148
0.601.890.0601.0009.2000.7823.816
0.652.160.0661.10010.4340.8074.294
0.702.310.0701.17010.8330.8804.760
V (m/s)uy (m)y /hy /y Fr Fr
0.450.320.0370.5681.780.7410.431
0.500.900.0400.6073.840.7981.900
0.550.900.0450.6954.000.8272.100
0.601.000.0460.6984.070.8922.022
0.650.700.0480.7364.300.9472.296
0.700.870.0470.7264.171.0302.356
V (m/s)uy (m)y /hy /y Fr Fr
0.451.500.0450.456.120.6773.109
0.501.600.0540.546.680.6863.230
0.551.800.0580.587.220.7293.439
0.602.000.0660.667.330.7453.808
0.652.200.0670.679.610.8014.401
0.702.300.0700.7010.760.84354.514
V (m/s)uy (m)y /hy /y Fr Fr
0.450.950.0360.365.000.7572.021
0.501.060.0380.385.040.8182.269
0.551.150.0430.435.150.8462.505
0.601.280.0430.434.780.9232.526
0.651.280.0430.434.671.0002.534
0.701.300.0510.514.640.9892.423
Geometryu (m/s)u (m/s)E (m)E (m)ΔE/E Efficiency (%)
Case 1 0.451.300.6620.1160.82482.46
0.501.500.6680.1460.78078.06
0.551.730.6730.1800.73173.19
0.601.890.6780.2070.69469.47
0.652.160.6870.4160.39339.36
0.702.310.6950.2950.57457.74
Case 20.450.320.6470.0610.90590.54
0.500.900.6520.0670.89589.59
0.550.900.6610.0660.89989.99
0.601.000.6640.0790.88088.09
0.650.70.6700.0500.92392.39
0.700.870.6720.0500.92492.42
Case 30.451.500.6550.1390.78678.68
0.501.600.6660.1550.76676.68
0.551.800.6730.1920.71471.46
0.602.000.6840.2300.66266.26
0.652.200.6880.2720.60360.39
0.702.300.6950.2950.57457.47
Case 40.450.950.6460.0650.89789.78
0.501.060.6500.0780.87987.97
0.551.150.6580.0870.86786.72
0.601.280.6610.1060.83883.89
0.651.280.6640.1070.83783.76
0.701.300.6750.1110.83583.55
The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

Farooq, U.; Li, S.; Yang, J. Numerical Analysis of Flow Characteristics and Energy Dissipation on Flat and Pooled Stepped Spillways. Water 2024 , 16 , 2600. https://doi.org/10.3390/w16182600

Farooq U, Li S, Yang J. Numerical Analysis of Flow Characteristics and Energy Dissipation on Flat and Pooled Stepped Spillways. Water . 2024; 16(18):2600. https://doi.org/10.3390/w16182600

Farooq, Umar, Shicheng Li, and James Yang. 2024. "Numerical Analysis of Flow Characteristics and Energy Dissipation on Flat and Pooled Stepped Spillways" Water 16, no. 18: 2600. https://doi.org/10.3390/w16182600

Article Metrics

Article access statistics, further information, mdpi initiatives, follow mdpi.

MDPI

Subscribe to receive issue release notifications and newsletters from MDPI journals

Existence and Stability of Nonmonotone Hydraulic Shocks for the Saint Venant Equations of Inclined Thin-Film Flow

  • Open access
  • Published: 13 September 2024
  • Volume 248 , article number  82 , ( 2024 )

Cite this article

You have full access to this open access article

froude experiment

  • Grégory Faye   ORCID: orcid.org/0000-0001-7390-4295 1 ,
  • L. Miguel Rodrigues 2 ,
  • Zhao Yang 3 &
  • Kevin Zumbrun 4  

Extending the work of Yang–Zumbrun for the hydrodynamically stable case of Froude number \(F<2\) , we categorize completely the existence and convective stability of hydraulic shock profiles of the Saint Venant equations of inclined thin film flow. Moreover, we confirm by numerical experiment that asymptotic dynamics for general Riemann data is given in the hydrodynamic instability regime by either stable hydraulic shock waves, or a pattern consisting of an invading roll wave front separated by a finite terminating Lax shock from a constant state at plus infinity. Notably, profiles, and existence and stability diagrams, are all rigorously obtained by mathematical analysis and explicit calculation.

Avoid common mistakes on your manuscript.

1 Introduction

In [ 37 , 39 ] there was carried out a comprehensive study of existence and nonlinear stability of hydraulic shock profiles for the Saint Venant equations (SV) of inclined thin film flow, under the assumption of hydrodynamic stability (or stability of constant solutions) of their endstates, a necessary condition for stability of shock profiles in standard Sobolev norms. It was shown under this condition that all profiles are monotone decreasing and nonlinearly stable . Notably, this conclusion includes both smooth and discontinuous (“subshock” containing) profiles.

In this paper, motivated by studies [ 30 , 31 ] of the closely related Richard–Gavrilyuk model (RG) for inclined thin film flow, in which nonmonotone profiles, and profiles with hydrodynamically unstable endstates, play a prominent role in asymptotic behavior, we revisit this problem in more detail, seeking nonmonotone profiles in the hydrodynamically unstable case. These of course cannot be stable in standard Sobolev norms, but as seen in [ 30 ], they can nonetheless be convectively stable , or stable in an appropriately exponentially-weighted norm, hence relevant to time-asymptotic behavior. Interestingly, we do find such waves, in a case that was neglected Footnote 1 in [ 39 ], and they appear to be convectively stable over a certain, computable regime.

The above observations have motivated the development of convective counterparts [ 15 , 16 ] to general results converting spectral stability into linear and nonlinear stability results [ 11 , 14 ]. Specializing [ 15 ] to the present case, we are able to supplement our complete spectral classification with corresponding nonlinear stability results.

The inviscid Saint-Venant equations in Eulerian, nondimensionalized form appear as

where h is fluid height; \(q=hu\) is total flow, with u fluid velocity; and \(F>0\) is the Froude number , a nondimensional parameter depending on reference height/velocity and inclination.

These form a \(2\times 2\) relaxation system , with associated formal equilibrium equation

where \(q_*\) is determined by the equilibrium condition that the second component of the right-hand side of ( 1.1 ) vanish. The first-order, principal part of ( 1.1 ), meanwhile, coincides with the equations of isentropic \(\gamma \) - law gas dynamics with \(\gamma =2\) [ 8 ]. System ( 1.1 ) admits constant solutions in the form of equilibria \((h,q)=(h_0,q_*(h_0))\) . Stability of constant solutions, known as hydrodynamic stability , is equivalent for \(2\times 2\) relaxation systems to the subcharacteristic condition that the equilibrium characteristic \(q_*'(h_0)\) of ( 1.2 ) lies between the characteristic speeds of ( 1.1 ), yielding the classical condition of Jeffreys [ 20 ],

Note the very special property that the condition does not depend on the particular value of \(h_0\) . For further discussion, see [ 21 , 39 ] on ( 1.1 ) and [ 2 , 3 , 4 , 32 ] on its viscous counterpart.

In the hydrodynamically stable regime \(F<2\) , one does expect persistent asymptotically-constant traveling wave solutions

analogous to shock waves of ( 1.2 ), known as relaxation shocks , or relaxation profiles, as verified in [ 39 ]. However, the hydrodynamically unstable regime \(F>2\) is also of interest, in both the convectively stable regime, since this is compatible with the description of large-time dynamics arising from compactly supported perturbations of Riemann data, and, in any case, as a scenario for complex behavior and pattern formation [ 4 , 30 ], with profiles ( 1.4 ) serving as potential building blocks for more complicated patterns. Here, we carry out an exhaustive study of existence and convective stability of hydraulic (SV) shocks for general F , including both cases \(F\gtrless 2\) .

1.1 Results

We now briefly state our main results, to be expanded in the remainder of the paper.

1.1.1 Existence

(Section  2 ) Expanding on the results of [ 39 ] for \(F<2\) , we categorize in Proposition 2.1 all possible types of possible hydraulic shocks: namely, the three monotone types (i), (iv), and (v) noted in [ 39 ], together with two new nonmonotone types (ii) and (iii) arising for \(F>2\) . These are displayed graphically in the left and right panels of Fig.  2 , the left one organized by the parameter \(H_R/H_L\) used in [ 39 ] and the right one by a new, more convenient parameter \(\nu _0:=\sqrt{H_{max}/H_{min}}\) in which the figures are more clear. Here \(H_L\) and \(H_R\) refer to the left and right limiting heights of the traveling wave, and \(H_{max}\) and \(H_{min}\) to the maximum and minimum heights. We note that type (ii)–(v) waves connect equilibria \((H_L,H_R)\) corresponding to shocks of the scalar equilibrium system ( 1.2 ), whereas type (i) waves are smooth, monotone increasing in height, and connect \((H_L,H_R)\) in the direction of a “reverse shock” of ( 1.2 ). The former, “forward-equilibrium shocks” exist precisely for

1.1.2 Spectral Stability

(Sections  3 and 4 ) In Sect.  3 , we investigate stability of essential spectra in the class of scalar weighted norms, or, equivalently here, stability of absolute spectra. We show that this fails for type (i) waves, corresponding to “reverse” equilibrium shocks, but is satisfied for type (ii)–(v) waves under condition

(always satisfied for cases (iii)–(v)) slightly stronger than the existence condition ( 1.5 ). Indeed, as noted in Remark 3.4 , essential stability fails for type (i) waves and for type (ii) waves failing to satisfy ( 1.6 ) in any type of weighted norm, yielding a conclusive result of convective instability for these types. We specifically refer to Sect.  3.5 for a complete summary of the results, which are also displayed graphically in the panels of Fig.  3 .

In Sect.  4 , we study stability of point spectrum for the remaining cases (ii), ( 1.6 ) and (iii)–(v), extending the generalized Sturm-Liouville argument introduced in [ 37 , 39 ] for the treatment of cases (iv)–(v). Remarkably, we are able to rigorously verify stability of point spectrum whenever the essential stability condition ( 1.6 ) is satisfied. Taken together, these results completely characterize spectral stability of hydraulic shocks of all types (see Proposition  4.3 ).

1.1.3 Linear and Nonlinear Stability

(Section  5 ) In Sect.  5 , we investigate for spectrally stable waves the questions of linear and nonlinear stability, providing a result of convective asymptotic time-exponential orbital stability, or convergence to a translate of the original traveling wave. This implies in particular, time-exponential stability under localized (e.g., Gaussian- or compact-support) perturbations, a result that is new even for the \(F<2\) case considered in [ 39 ]. We have chosen here to derive these results by specializing to ( 1.1 ) the general theory from [ 15 ]. Despite the fact that analyzing directly ( 1.1 ) would come with significant simplifications due to the special structure of systems of two equations compared to general systems, a detailed analysis would still be rather technical and long, without conveying much specific insight about the dynamics at hand. We stress moreover that, though the results of [ 39 ] do not apply to the present case, their proof does contain all the main ingredients to yield the nonlinear stability of interest. Again, though a simpler form of the arguments of [ 39 ] would be sufficient here, since time-exponential decay is simpler to handle than time-algebraic decay, a self-contained exposition of such an adaptation would still be rather long and technical.

1.1.4 Global Time-Asymptotic Dynamics

(Section  6 ) Finally, in Sect.  6 , we carry out using CLAWPACK [ 9 , 25 ] numerical experiments with (perturbed and unperturbed) “Riemann” or “dambreak” data consisting of constant equilibrium states to either side of an initial jump discontinuity, testing the “real life” validity of our rigorous existence/stability conclusions, in the sense of large-amplitude perturbations and resulting time-asymptotic behavior, or “generalized Riemann solution”. We see that our analytically derived stability conditions indeed predict not only small-perturbation stability or instability, but large-scale asymptotic behavior. Specifically, when stability holds, the asymptotic response to even large-scale localized perturbations is convergence to a hydraulic shock, monotone or nonmonotone as the case may be.

When stability fails, on the other hand—recall, through instability of essential spectrum , having to do with convective stability of the constant left endstate of the shock—we see bifurcation to an “invading front” connecting roll wave patterns on the left to a constant state on the right: that is, an “essential bifurcation” such has been studied for smooth waves of reaction-diffusion systems in [ 34 ]. Thus, our (local) stability conditions indeed successfully predict large-scale asymptotic behavior. Interestingly enough, in all our experiments the expanding speed of the instability pattern is well-predicted by the heuristics of [ 13 ].

1.2 Discussion and Open Problems

The Saint Venant model has proven remarkably amenable to analysis, admitting complete solutions to both existence and stability questions now in a variety of settings. The present analysis fits among this list, giving complete and definitive answers to the questions of existence and convective stability of hydraulic shock solutions. In particular, the fact that absolute and point spectral stability could be completely characterized is quite remarkable and apparently special to (SV). It is a very interesting open problem to what extent the Sturm-Liouville arguments used here might extend to large-amplitude traveling waves of general \(2\times 2\) relaxation systems under a condition of convectively stable essential spectrum, generalizing the treatment by Liu [ 24 ] of small-amplitude waves in the hydrodynamically stable case.

The analyses of linear and nonlinear stability in [ 39 ] also rely in places on specific computations for (SV). However, different from the situation as regards spectral stability, the strategies for converting spectral to nonlinear stability are rather general, and could be expected under appropriate structural conditions to carry over to the general case of relaxation models. These considerations motivate a more general and systematic study of such problems, as done by the first two authors for exponentially spectrally stable Riemann shocks [ 14 ], and will be the object of a future publication [ 15 ] from which we already borrow some results.

Jointly with [ 21 , 37 , 39 ], the present contribution provides for (SV) an almost complete classification of traveling waves from the point of view of existence and spectral stability. Nevertheless we would like to point out that even in (SV), at the spectral level, a stability classification of waves that have characteristic points but are not periodic is still missing. Likewise one of the outstanding remaining puzzles in the nonlinear stability of relaxation waves, either smooth or discontinuous, is the treatment of waves with characteristic points, generalizing to the system case the scalar analysis of [ 12 ]. At the nonlinear level, the corresponding difficulties are expected to occur also in the analysis of the dynamics near roll waves, which has not been touched even for (SV); see for example the discussions of [ 21 ]. Indeed, there are some additional difficulties for (SV) due to an infinite-dimensional center manifold coming from degeneracy of the model [ 21 ]. We find this to be the main open problem in the theory of general (including periodic) traveling waves.

Finally, we mention as an interesting direction for further investigation, the rigorous treatment of the phenomenon of essential bifurcation/invading roll wave fronts that we see in our numerical experiments, the lack of smoothness and parabolic smoothing making this a nonstandard problem not covered by the methods of [ 34 ] and related references.

2 Existence of Traveling Waves

In this section, we recover the basic existence theory from [ 39 , Prop. 1.1], in the process unraveling the nonmonotone case omitted there. We focus on traveling waves with piecewise smooth profiles without characteristic point, neither on profiles nor at infinity. The presence of characteristic points is expected to have dramatic effects on the existence, spectral stability and nonlinear dynamics; see the related analyses in [ 12 , 21 ]. We also restrict to waves with nonnegative velocities so that absolute values may be dropped, but one may be careful to check as a consistency condition that indeed \(Q\ge 0\) .

To expect some form of uniqueness when dealing with discontinuous solutions we need to impose some form of entropy conditions. Combined with the non-characteristic assumption, even the weaker forms of the latter imply that the traveling wave profiles exhibit at most one discontinuity. We again refer to [ 12 , 21 ] for a detailed discussion. Without loss of generality, by translational invariance, the discontinuity of wave profiles may be fixed at \(x=0\) . When restricting further to asymptotically constant profiles, they also yield that limiting endstates are distinct.

Here and elsewhere, let \([h]_x:=h(x^+)-h(x^-)\) of a quantity h across a discontinuity located at x , and \([h]:=[h]_0\) . In smooth regions, traveling-wave solutions ( 1.4 ) satisfy

whereas at discontinuities, we have the Rankine–Hugoniot conditions

A simple observation is that the end states \((H_R,Q_R)\) and \((H_L,Q_L)\) of the traveling wave profiles ( 1.4 ) must be equilibria, that is \(Q_{L,R}=q_*(H_{L,R})=H_{L,R}^{3/2}\) (since we are working in the physical range \(H>0\) ). Combined together the first halves of ( 2.1 ) and ( 2.2 ) are equivalent to the existence of a constant \(q_0\) such that

With such a \(q_0\) fixed, the second equation of ( 2.1 ) leads to the scalar ODE

while the second Rankine–Hugoniot condition in ( 2.2 ) reads

Equation ( 2.4 ) is a scalar first-order ODE, so that it cannot connect smoothly an endstate to itself (in a non stationary way). We have already discussed that when instead a discontinuity is indeed present, we must have \((H_R,Q_R)\ne (H_L,Q_L)\) so that in any case \(H_L\ne H_R\) . Therefore from ( 2.3 ) stems that \((H_L,H_R, c,q_0)\) must satisfy

and necessarily \(c>0\) , \(q_0>0\) . Note that then the condition \(Q\ge 0\) becomes \(H\ge q_0/c\) with

Moreover, from the sign of \(q_0\) and entropy conditions stem, when a discontinuity is present,

which is equivalently written as

The scalar ODE ( 2.4 ) may be factorized as

Note that in any case \(H_{out}<q_0/c<\min (\{H_L,H_R\})\) and recall that solutions to ( 2.6 ) taking values below \(q_0/c\) have no significance for the original traveling wave profile problem. Therefore for the discontinuous profiles, one needs \(H_{out}<H_R<H_s<H_L\) and a simple one-dimensional phase-portrait analysis shows that the piece converging to \(H_R\) must be constant. As a consequence, in terms of \(H_*=H(0^-)\) , ( 2.5 ) is reduced to

which possesses a unique positive solution distinct from \(H_R\)

Note that \(H_s>H_R\) implies \(H_*>H_R\) .

For the sake of comparison with [ 39 ], let us introduce the scaling parameter

and express the above quantities as

We have the following extension/correction of [ 39 , Prop. 1.1] cases (ii) and (iii) were mistakenly omitted there:

Proposition 2.1

Let \((H_L,H_R)\) be a couple of positive heights.

When \(H_L<H_R\) , that is when \(\nu <1\) , there exists only one kind of non-characteristic wave profiles connecting \(H_L\) to \(H_R\) ,

increasing smooth profiles , that do exist if and only if \(H_L<H_R<H_s\) , that is, if and only if

When \(H_R<H_L\) , that is when \(\nu >1\) , there exist four kinds of non-characteristic waves connecting \(H_L\) to \(H_R\) ,

nonmonotone discontinuous profiles , consisting of a smooth portion increasing from \(H_L\) to \(H_*\) , connected by an entropy-admissible Lax shock to a portion constant equal to \(H_R\) , that do exist if and only if \(H_R<H_s<H_L<H_*\) , that is, if and only if

Riemann profiles , consisting of a portion equal to \(H_L\) , connected by an entropy-admissible Lax shock to a portion constant equal to \(H_R\) , that do exist if and only if \(H_R<H_s<H_*=H_L\) , that is, if and only if

decreasing discontinuous profiles , consisting of a smooth portion decreasing from \(H_L\) to \(H_*\) , connected by an entropy-admissible Lax shock to a portion constant equal to \(H_R\) , that do exist if and only if \(H_R<H_s<H_*<H_L\) , that is, if and only if

smooth decreasing profiles , that do exist if and only if \(H_s<H_R<H_L\) , that is, if and only if

Simple one-dimensional phase-portrait considerations provide the classification in terms of respective positions of \(H_L\) , \(H_R\) , \(H_s\) and \(H_*\) , that may be readily translated as conditions on \(\nu \) and F . It only remains to point out that, in case (ii), we have used that, when \(\nu >1\) ,

to discard as redundant one of the inequalities. Incidentally we also point out that when \(\nu >1\)

so that case (ii) is indeed non empty. This completes the proof. We refer to Fig. 1 for an illustration of the various hydraulic shock profiles. \(\square \)

figure 1

Examples of cases (i)–(v) of hydraulic shock profiles prescribed in Proposition  2.1 . (i) an increasing smooth profile with \(H_L=1\) , \(H_R=1.3\) , \(F=3\) ; (ii) a nonmonotone discontinuous profile with \(H_L=1\) , \(H_R=0.4\) , \(F=3\) ; (iii) a Riemann profile with \(H_L=1\) , \(H_R=0.4\) , \(F=\tfrac{\sqrt{7}}{2}+\sqrt{\tfrac{7}{10}}\) ; (iv) a decreasing discontinuous profile with \(H_L=1\) , \(H_R=0.2\) , \(F=1.5\) ; (v) a smooth decreasing profile with \(H_L=1\) , \(H_R=0.8\) , \(F=1.5\)

With hydrodynamical stability in mind, let us compare different \(\nu \) -dependent Froude thresholds to the critical value 2. For any \(\nu <1\) ,

so that case (i) is always hydrodynamically unstable. When \(\nu >1\) ,

the latter inequality following from the fact that its left-hand side is increasing with \(\nu \) and takes the value 2 at \(\nu =1\) . Thus cases (ii) and (iii) are always hydrodynamically unstable, case (v) is always hydrodynamically stable and case (iv) may or may not be hydrodynamically unstable. Case (v), and case (iv) when \(F<2\) have been thoroughly analyzed in [ 37 , 39 ].

In the above discussion, we have decided in advance that we were looking for non-characteristic traveling waves connecting \(H_L\) to \(H_R\) . For the convenience of the reader, we now provide a more systematic treatment of non constant waves in terms of

with \((H_{min},H_{max})\) now merely playing the role of wave parameters (replacing \((c,q_0)\) ).

When \(F>\nu _0(\nu _0+1)\) , only waves of case (i) exist, with \(H_L=H_{min}\) and \(H_R=H_{max}\) .

When \(F=\nu _0(\nu _0+1)\) , \(H_s=H_{max}\) and there exist two families of traveling waves, one family with each member beginning by a smooth infinite portion arising from \(H_L=H_{min}\) , connected by a Lax shock to an infinite array of increasing portions passing though \(H_s\) , connected by Lax shocks, the family being parameterized by an arbitrary sequence of lengths taken in \((0,+\infty )^{{\mathbb {N}}}\) ; the other family with each member consisting in an infinite Footnote 2 array of increasing portions passing though \(H_s\) , connected by Lax shocks, the family being parameterized by an arbitrary sequence of lengths taken in \((0,+\infty )^{{\mathbb {Z}}}\) . The latter family includes periodic “roll wave” solutions of the type discovered by Dressler [ 10 ], that is, periodic traveling-wave solutions with exactly one discontinuity and one characteristic point by period. A comprehensive study of their spectral stability may be found in [ 21 ].

only waves of cases (ii)–(iii) and (iv) exist, with \(H_R=H_{min}\) and \(H_L=H_{max}\) .

there exists no wave.

only waves of case (v) exist, with \(H_R=H_{min}\) and \(H_L=H_{max}\) .

We summarize the existence results in Fig.  2 .

figure 2

Left panel: domains of cases (i)–(v) from Proposition  2.1 , extending the scope of [ 39 , Fig. 3. (b)] beyond the box \(0<H_R/H_L<1\) , \(0<F<2\) (note that, by re-scaling, \(H_L\) is fixed to be 1 in [ 39 ]); Right panel: visualization of domains of cases (i)–(v) (2) (4) by incorporating \(\nu _0\) from Remark  2.3

3 Spectral Framework and Essential Spectrum

We now turn to an examination of the spectral stability of waves listed in Proposition  2.1 . When doing so, we use extensively standard elements of spectral theory specialized to nonlinear wave stability. We give little detail on those but rather refer the reader to the already classical [ 22 , 26 , 33 , 40 , 41 ] for detailed comprehensive exposition and to the recent [ 7 ] for a self-contained worked-out case that could hopefully be used as a gentle entering gate. For discontinuous waves, this involves, at least implicitly, Evans-Lopatinskiĭ determinants, that interpolate between pure Evans functions used in smooth wave analysis and pure Lopatinskiĭ determinants used to analyze local-in-time persistence near shocks. On the latter we refer for instance to [ 6 , Section 4.6]. Evans-Lopatinskiĭ determinants are commonly encountered in the literature about spectral and linear stability of shocks; see for instance [ 17 , 18 , 21 , 38 ].

3.1 Linearization and Spectrum

To introduce the relevant spectral problem in a concise way, let us write ( 1.1 ) in standard abstract form

System ( 3.1 ) must be satisfied at least in weak sense, thus, for piecewise smooth solutions we impose ( 3.1 ) to hold in a strong sense on domains corresponding to smooth parts and along a jump whose location at time t is at \(\varphi (t)\) we impose Rankine–Hugoniot jump conditions

Pick a non-characteristic traveling wave of profile \(W:=(H,Q)\) and speed c . When W is smooth, writing equations in terms of v , with \(w(t,x)=W(x-ct)+v(t,x-ct)\) , and replacing nonlinear terms with a source term, lead to

where the source term F depends on space and time but the matrix-valued coefficients A and E depend only on x and are explicitly given by

In turn, when W possesses a discontinuity at 0, proceeding in the same way but in terms of \((v,\psi )\) , with \(w(t,x)=W(x-(ct+\psi (t)))+\tilde{w}(t,x-(ct+\psi (t)))\) , \(v=\tilde{w}-\psi \,W'\) , yields

where ( A ,  E ,  F ) are as above, G is a time-dependent source term and \({\mathbb {R}}^*:={\mathbb {R}}\backslash \{0\}\) .

It is customary in smooth wave analysis to directly discard source terms. This is justified by the fact that when considering initial value problems one may recover the general source-term case through Duhamel’s formula. However, for discontinuous waves, the linearized problem is a mixed initial boundary value problem and the arguments fails. The source terms G that may be recovered by the Duhamel formula are those that are pointwise in time colinear with [ W ]. On a directly related note, let us observe that whereas ( 3.3 ) directly fits in standard semigroup theory, ( 3.4 ) does not but it does belong to the class of problems that can be analyzed through the more general, infinite-dimensional Laplace transform theory, as covered in [ 1 ], and we shall extrapolate standard spectral terminology to this case.

Applying the Laplace transform to the above linearized problems yields respectively

with a different meaning for \((v,\psi ,F,G)\) , and \(\lambda \in {\mathbb {C}}\) a spectral parameter. For the sake of concision, let us set

in respective cases.

For some choice of functional spaces ( X ,  Y ), we say that \(\lambda \) does not belong to the ( X ,  Y )-spectrum of either ( 3.3 ) or ( 3.4 ) if and only if \(L_\lambda \) is invertible as a bounded operator from Y to X . In the smooth case, this matches the classical definition of the spectrum of the generator of the dynamics on X , when Y is chosen to be the corresponding domain.

Consistently we say that the wave under consideration is spectrally ( X ,  Y )-stable if the corresponding ( X ,  Y )-spectrum is included in \(\{\lambda \in {\mathbb {C}};\textrm{Re}(\lambda )\le 0\}\) and that it is spectrally ( X ,  Y )-unstable otherwise. We say that it is exponentially spectrally ( X ,  Y )-stable if there exists \(\theta >0\) such that the ( X ,  Y )-spectrum is included in \(\{\lambda \in {\mathbb {C}};\textrm{Re}(\lambda )\le -\theta \}\cup \{0\}\) and 0 has multiplicity 0 if the wave is smooth and \(W'\notin Y\) , 1 otherwise.

When stability/instability is considered with respect to \((X,Y)=(L^2({\mathbb {R}};{\mathbb {C}}^2), H^1({\mathbb {R}};{\mathbb {C}}^2))\) in the smooth case, or \((X,Y)=(L^2({\mathbb {R}}^*;{\mathbb {C}}^2)\times {\mathbb {C}},H^1({\mathbb {R}}^*;{\mathbb {C}}^2)\times {\mathbb {C}})\) in the discontinuous case, we drop any mention to the functional pair ( X ,  Y ). This particular choice of functional spaces takes into account that our profiles are non-characteristic. From this property also stems that the spectrum is not really affected by the level of regularity encoded by functional spaces provided that they are chosen consistently. However it is strongly impacted by the level of localization.

To take this into account, we introduce for \((\eta _L,\eta _R)\in {\mathbb {R}}^2\) the weighted spaces

with \(\mathcal {X}=L^2\) or \(\mathcal {X}=H^1\) . Consistently, when talking about stability, we replace any mention to a pair ( X ,  Y ) with the adverb convectively if it can be achieved respectively with \((X,Y)=(L^2_{\eta _L,\eta _R}({\mathbb {R}};{\mathbb {C}}^2),H^1_{\eta _L,\eta _R}({\mathbb {R}};{\mathbb {C}}^2))\) or \((X,Y)=(L^2_{\eta _L,\eta _R}({\mathbb {R}}^*;{\mathbb {C}}^2)\times {\mathbb {C}},H^1_{\eta _L,\eta _R}({\mathbb {R}}^*;{\mathbb {C}}^2)\times {\mathbb {C}})\) for some \((\eta _L,\eta _R)\) such that \(\eta _L\ge 0\) and \(\eta _R\le 0\) . Correspondingly convective instability refers to the failure of convective stability. When it will be convenient to keep track of the chosen weights we will replace the general term “convectively” with the more specific term “ \((\eta _L,\eta _R)\) -weightedly”.

The constraint ( \(\eta _L\ge 0\) and \(\eta _R\le 0\) ) imposed in the definition of convective stability is motivated by the will to pave the way for nonlinear analysis. At a semi-abstract level, a functional space Z appearing at the spectral level (for scalar components) is thought as a good space for nonlinear analysis if \(Z\cap L^\infty \) is an algebra. This leads to the above requirements on weights. In the discontinuous case, another obstruction to a nonlinear analysis may be anticipated. Indeed in a Duhamel formulation source terms would contain terms that decay spatially like the square of components of \(\psi \,W'\) , which may belong to a \((\eta _L,\eta _R)\) -weighted space only if Footnote 3 \(W'\equiv 0\) or

The situation is dramatically different in the smooth case since there one needs to introduce a phase shift (which would also appear in nonlinear terms) only if \(W'\) does belong to the kernel of \(L_0\) , that is, only if \(\eta _L<\eta _L^\infty \) and \(\eta _R>\eta _R^\infty \) where \(\eta _L^\infty \) is as above and

which do imply \(\eta _L<2\eta _L^\infty \) and \(\eta _R>2\eta _R^\infty \) . In the discontinuous case, a phase shift is required no matter what; in the foregoing derivation of ( 3.4 ) we have partially hidden it when we have moved from \(\tilde{w}\) to v . In our definition, for the sake of simplicity, we have chosen not to include the extra constraint \(\eta _L<2\eta _L^\infty \) of the discontinuous case but as we check in Remark  3.7 it turns out that in the present case extra constraints already enforce \(\eta _L<\eta _L^\infty \) .

Our current definition of convective stability/instability uses scalar exponential weights. Though this choice is the most usual one, it is also somewhat arbitrary. However, as we shall detail in Remark  3.4 , in the present case, no substantial further gain in stabilization may be expected from the use of more complex weights.

3.2 Essential Spectrum, Consistent Splitting and Absolute Instability

A subset of the ( X ,  Y )-spectrum is constituted of the \(\lambda \) such that \(L_\lambda \) is not Fredholm of index 0 as a bounded operator from Y to X . By analogy with the standard case, we call this part the ( X ,  Y )-essential spectrum. The essential spectrum is therefore the set of \(\lambda \) such that the codimension of the range of \(L_\lambda \) and the dimension of its kernel are not equal, a clear obstruction to invertibility, which occurs when both are zero.

By using that being Fredholm of index 0 is invariant by compact perturbations and that the problem at hand is non characteristic with coefficients converging exponentially fast to their limits, one may derive a characterization of the essential spectrum. We do not provide details on the proof of the latter but we refer the reader to the Appendix to [ 19 , Chapter 5] for a worked out version in a closely related context.

To discuss the outcome, we introduce

and recall that

Then \(\lambda \) does not belong to the \((\eta _L,\eta _R)\) -weighted essential spectrum if and only if \(G_{H_L}(\lambda )\) has no eigenvalue with real part \(\eta _L\) , \(G_{H_R}(\lambda )\) has no eigenvalue with real part \(\eta _R\) and the sum of the number of eigenvalues of \(G_{H_L}(\lambda )\) with real part greater than \(\eta _L\) and of the number of eigenvalues of \(G_{H_R}(\lambda )\) with real part lesser than \(\eta _R\) equals 2 in the smooth case, 1 in the discontinuous case.

By continuity in \(\lambda \) , \((\eta _L,\eta _R)\) -weighted stability requires that each of the above-mentioned numbers is constant in \(\lambda \) on \(\{\,\lambda ;\,\textrm{Re}(\lambda )>0\,\}\) , a property referred to as consistent splitting in part of the literature. Now, note that when \(|\lambda |\rightarrow \infty \) , eigenvalues of \(G_h(\lambda )\) expand as

which when specialized to \(h=H_L\) or \(H_R\) is equivalently written as

The leading order part of these spatial eigenvalues is given by the eigenvalues of \(-\lambda \,A_h^{-1}\) and thus is directly connected to the characteristic velocities of \(\partial _t+A_h\,\partial _x\) . As a consequence, for \(h=H_L\) or \(H_R\) , for any \(\eta >0\) there exists \(C_\eta >0\) such that when \(|\lambda |\ge C_\eta \) and \(\textrm{Re}(\lambda )\ge \eta \) , \(G_{H_L}(\lambda )\) has two eigenvalues with positive real parts when \(h<H_s\) and eigenvalues with real parts of opposite sign when \(h>H_s\) .

As a consequence, a specific way in which failure of convective stability (resp. absolute convective instability) may occur in the present case is when for \(h=H_L\) or \(H_R\) such that \(h>H_s\) , there exists \(\lambda \) with positive real part (resp. nonzero with nonnegative real part) such that the eigenvalues of \(G_h(\lambda )\) have the same real part. This scenario matches what is commonly designated in the literature as failure of extended consistent splitting or absolute instability . To decide whether an absolute instability may indeed occur, let us first make explicit that the eigenvalues of \(G_h(\lambda )\) are given as

for some determination of \(\sqrt{{{\mathcal {Q}}}_h(\lambda )}\) . Note that \(\gamma _{\pm ,h}(\lambda )\) share the same real part exactly when \({{\mathcal {Q}}}_h(\lambda )\) is a nonpositive real number. Since

one readily deduces that the latter does occur for some \(\lambda \) with positive real part (resp. nonzero with nonnegative real part) if and only if

In the present case, when convective stability (resp. exponential convective stability) fails in the foregoing way, there also exists a \(\lambda \) with positive real part (resp. nonzero with nonnegative real part) such that \(\gamma _{\pm ,h}(\lambda )\) are equal. At this \(\lambda \) , the resolvent operator cannot be continuously extended even as an operator from the space of test functions to distributions. This shows that, in the present case, such absolute instabilities cannot be cured in any sensible sense, in particular not by replacing exponential weights by a more general class of reasonable weights.

In order to elucidate further a possible absolute instability, we compute that when \(h=H_L\) or \(h=H_R\) ,

Recalling that the scenario also requires \(h>H_s\) and observing that, when \(\nu >1\) ,

one deduces that absolute instability may only occur in case (ii) of Proposition  2.1 and does occur when

3.3 Smooth Fronts

We temporarily restrict the discussion to smooth profiles, that is, to cases (i) and (v) of Proposition  2.1 .

Case (v) has already been studied in [ 37 , Section 3] with conclusion that all profiles of case (v) are spectrally stable but not exponentially spectrally stable. With a few more simple computations one may even check that this spectral stability is of diffusive type in a sense compatible with the application of general results from [ 27 ] and conclude to nonlinear asymptotic stability with algebraic decay rates.

The only question left concerning case (v) is whether convective exponential spectral stability also holds. In this case the only obstacle to exponential spectral stability without weight is the presence of two curves of essential spectrum passing through \(\lambda =0\) tangentially to the imaginary axis, one for each spatial infinity. Recall that since, in case (v), \(H_R>H_s\) and \(H_L>H_s\) , there holds for \(h=H_L\) , \(H_R\) ,

Therefore, to conclude convective exponential spectral stability, one needs only to check that curves of essential spectrum near \(\lambda =0\) are due to changes of sign of \(\textrm{Re}(\gamma _{+,H_L}(\lambda ))\) and \(\textrm{Re}(\gamma _{-,H_R}(\lambda ))\) . Since

one concludes convective exponential spectral stability in case (v) from the fact that when \(\nu >1\)

In turn, in case (i), \(H_R<H_s\) and \(H_L<H_s\) so that for \(h=H_L\) , \(H_R\) ,

Therefore to prove that convective spectral stability fails it is sufficient to prove that a spectral instability is caused by what happens near \(-\infty \) , thus with \(h=H_L\) . This follows from the fact that in case (i), \(F>2\) , and hence both endstates generate an essential spectrum instability.

3.4 Discontinuous Fronts

We now specialize to discontinuous fronts, as in cases (ii), (iii) and (iv) of Proposition   2.1 . Our goal in the present section is to completely elucidate the effect of essential spectrum on stability/instability of any type so as to reduce the issues to the examination of unstable eigenvalues, carried out in the next section.

For all the cases under consideration here, \(H_R<H_s<H_L\) thus

This readily implies that instabilities due to the behavior near \(+\infty \) may always be convectively stabilized whereas the convective stabilization of instabilities due to the behavior near \(-\infty \) require that those occur through a change of sign in \(\textrm{Re}(\gamma _{+,H_L}(\lambda ))\) . Note that at this stage it is not clear whether the latter necessary condition is also sufficient.

In order to decide this necessary condition, we compute that

As a consequence,

In particular, the condition is at least met in the high-frequency regime. Another necessary condition is that \(\textrm{Re}(\lambda )>0\) implies \(\textrm{Re}(\gamma _{+,H_L}(\lambda ))<\textrm{Re}(\gamma _{-,H_L}(\lambda ))\) . We have already examined the latter condition when discussing absolute instability and proved that it fails only in case (ii) when

Moreover when \(F<\sqrt{2\nu (\nu +1)}\) , \(\textrm{Re}(\lambda )\ge 0\) also implies \(\textrm{Re}(\gamma _{+,H_L}(\lambda ))<\textrm{Re}(\gamma _{-,H_L}(\lambda ))\) .

The full condition we want to elucidate is

With explicit expressions ( 3.7 ) in mind, we first recall that

and observe that in present cases, when \(\textrm{Re}(\lambda )\ge 0\) ,

This motivates the following lemma:

For any positive \(\alpha \) , \(\gamma \) ,

From the classical formula \((\textrm{Re}(\sqrt{z}))^2=(\textrm{Re}(z)+|z|)/2\) , we deduce

Direct computations yield that

and that either \(\Gamma '\) is constantly zero, which happens when \(\beta ^2=4\alpha \gamma \) , or that it never vanishes, when \(\beta ^2\ne 4\alpha \gamma \) . Hence the result by monotonicity. \(\square \)

When applying the lemma to an estimate of \(\textrm{Re}({{\mathcal {Q}}}_h(\lambda ))\) , we want to determine what is the minimum obtained from the lemma. This stems from the computation

Note that the latter expression does not depend on \(\lambda \) and that its sign is determined by the sign of \(F-2\) . All together we deduce that, to determine the convective stabilitization of the essential spectrum, when \(F\ge 2\) it is sufficient to discuss what happens at the limit \(\textrm{Im}(\lambda )\rightarrow \infty \) whereas when \(F<2\) it is sufficient to look at the case when \(\lambda \in {\mathbb {R}}\) .

As for the smooth profiles, the case \(F<2\) has been thoroughly analyzed in [ 37 ] and the only thing left is to check that one may also obtain exponential convective stability in this case. This follows from the same computation as for discontinuous profiles.

From now on we focus on discontinuous profiles when \(F\ge 2\) . In this context it follows from the previous lemma and the above \(|\lambda |\rightarrow \infty \) asymptotics that failure of convective stability by essential spectrum is equivalent to

This coincides with the condition for absolute instability

Let us emphasize that the coincidence of the boundary of absolute instability with the boundary of convective stability Footnote 4 defined through scalar exponential weights is not a general fact but a specific property of the present problem. It comes with the strong consequence that there is no need to consider more general weights. One may obtain a simple (but artificial) counterexample to a more general claim in this direction by simply considering as a single system two uncoupled systems requiring incompatible weights.

3.5 Summary

figure 3

The preceding analysis motivates the definition of the following regions in parameters space. We set

froude experiment

together with

froude experiment

We refer to Fig.  3 for a visualization of these regions in parameters space. So far, our results can be summarized as:

3.6 Maximal Decay Rate: Another View

Before moving on with the rest of the program, we would like to halt and offer a different perspective on the former computations so as to address the following question: what is the maximal essential spectral gap that may be opened by tuning our weights appropriately ? Since the boundaries of the essential spectrum due to what happens near \(+\infty \) may be pushed arbitrarily to the left of the compex plane, we may again focus on the contribution from the left. The same computation we have carried out to determine absolute instability yields as an upper bound for the essential spectral gap

and that it is reached with spatial decay rate

Reciprocally one may check with arguments similar to the ones used above (mostly relying on Lemma  3.5 ) that choosing \(\eta _L=\eta _{opt}\) and \(\eta _R\) sufficiently negative provides the optimal spectral gap.

The effect of moving \(\eta _L\) is illustrated in Fig.  4 . There the curves are obtained by solving in \(\lambda \in {\mathbb {C}}\) the equations

with parameter \(\xi \in {\mathbb {R}}\) as

and we have introduced

To prove the last claim in Remark  3.2 , we observe that

is indeed positive in the cases under consideration. To carry out the above computation, we have used that

figure 4

Visualization of some of the boundaries of the weighted essential spectrum as \(\eta _L\) is varied. In all figures, the blue curves represent the computed boundaries while the red half-lines represent the absolute spectrum which terminates at branch points marked by black crosses. Without weight, the essential spectrum is always unstable, while it is strictly stabilized with a gap for \(\eta _L\in \left( \eta _L^\textrm{min},\eta _L^\textrm{max}\right) \) . At the critical weights \(\eta _L=\eta _L^\textrm{min}\) or \(\eta _L=\eta _L^\textrm{max}\) , the weighted essential spectrum is only marginally stabilized with no spectral gap. The parameters are fixed to \((H_L,H_R,F)=(1,1/4,3)\) so that \((\eta _L^\textrm{min},\eta _L^\textrm{max})=(3,5)\)

4 Sturm–Liouville Analysis

Throughout this section, we consider a discontinuous profile and assume that F satisfies

We also fix some \(\eta _L\in \left( \eta _L^\textrm{min},\eta _L^\textrm{max}\right) \) where \(\eta _L^\textrm{min}\) and \(\eta _L^\textrm{max}\) are as above.

Here, as announced, we study possible unstable eigenvalues and, to do so, adapt the arguments from [ 37 ].

4.1 The Reduced Eigenvalue Problem

By imposing a vanishing on \({\mathbb {R}}_+\) , as we can for our purposes, we reduce the eigenfuction problem to finding a nonzero \((v,\psi )\) in Footnote 5 \(H^1_{\eta _L}({\mathbb {R}}_-^*;{\mathbb {C}}^2)\times {\mathbb {C}}\) solving

We begin by inspecting the special case when v is zero (but \(\psi \) is not). A direct inspection shows that it only happens when \(\left[ \,\lambda \,W-r(W)\right] \) is zero, which is equivalent to \(\lambda \) and [ r ( W )] both being zero (since the first component of [ r ( W )] is zero and the first of [ W ] is nonzero). The latter occurs exactly when we are in the Riemann shock case, case (iii) of Proposition  2.1 . Actually the vanishing of \(\left[ \,\lambda \,W-r(W)\right] \) when \(\lambda =0\) alters many of the considerations to come. For this reason we postpone the treatment of the Riemann shock case to the end of the present section.

Since we are now excluding the Riemann shock case, \(\left[ \,\lambda \,W-r(W)\right] \) is non zero and one may eliminate \(\psi \) to reduce the discussion further to the existence of a nonzero \(v=(v_1,v_2)\) in \(H^1_{\eta _L}({\mathbb {R}}_-^*;{\mathbb {C}}^2)\) such that on \({\mathbb {R}}_-^*\)

For the sake of writing simplification we introduce one flux coordinate and replace v with \(u=(u_1,u_2):=(v_1,-cv_1+v_2)\) . With this change, we turn the problem into finding a nonzero u in \(H^1_{\eta _L}({\mathbb {R}}_-^*;{\mathbb {C}}^2)\) such that on \({\mathbb {R}}_-^*\)

In the foregoing we have denoted by a the characteristic determinant

Let us now examine the possibility to have a nonzero solution u with zero component \(u_2\) . A direct inspection shows that this may happen only when \(\lambda =0\) and that the corresponding u is necessarily a multiple of \((H',0)\) . Note that reciprocally one checks readily that when \(\lambda =0\) necessarily \(u_2\equiv 0\) . Thus this situation corresponds exactly to the possibility of 0 being in the spectrum due to translational invariance.

We now focus on the case when \(u_2\) is not zero. Then the eigenvalue problem may be recasted into the problem of finding a nonzero \(u_2\) in \(H^2_{\eta _L}({\mathbb {R}}_-^*;{\mathbb {C}})\) solving

on \({\mathbb {R}}_-^*\) and

In order to match notation from [ 37 ] we introduce

so that the equation on \({\mathbb {R}}_-^*\) becomes

We point out for later use that from the fact that \((H',0)\) solves the interior ODE problem for \((u_1,u_2)\) when \(\lambda =0\) , one deduces that \(H''=-f_2\,H'\) .

At last, in order to symmetrize the interior equation, we perform a Liouville-type transformation and replace \(u_2\) with w defined by Footnote 6

This replaces the equation on \({\mathbb {R}}_-^*\) with

also written as

which is exactly [ 37 , Equation (2.14)], whereas the boundary condition becomes

Before going on we need to check that \(u_2\in H^2_{\eta _L}({\mathbb {R}}_-^*;{\mathbb {C}})\) implies \(w\in H^2({\mathbb {R}}_-^*;{\mathbb {C}})\) . From the analysis of the previous section we know that, when \(u_2\in H^2_{\eta _L}({\mathbb {R}}_-^*;{\mathbb {C}})\) , its spatial decay rate is precisely \(\textrm{Re}(\gamma _{-,H_L}(\lambda ))\) . Therefore this amounts to proving that

A direct computation shows that this is equivalent to

thus to the fact that \(\lambda \) does not belong to the absolute spectrum.

Therefore it is indeed sufficient to discard the possibility of a nonzero w in \(H^2({\mathbb {R}}_-^*;{\mathbb {C}})\) solving ( 4.1 )–( 4.2 ).

4.2 Non Real Growth Rates

We stress that whereas the interior part, ( 4.1 ), is symmetric on functions compactly supported in \({\mathbb {R}}_-^*\) , completing it with boundary condition ( 4.2 ) does not yield a symmetric operator. An argument, specialized to the case at hand, is thus needed to show that necessarily \(\lambda \in {\mathbb {R}}\) if such a w exists. We provide such a concrete argument now.

To begin with, we observe that, combined with ( 4.2 ), multiplying ( 4.1 ) with \(\overline{w}\) and integrating yield

When \(\textrm{Im}(\lambda )\ne 0\) , the imaginary part of ( 4.3 ) gives

Since \(c_1<0\) and \(f_3<0\) , the last equality implies

As a consequence, we need to study the sign of \(f_4-\frac{1}{2}f_1f_2-\frac{1}{2}f_1'\) . To determine this sign we observe that

Let us denote \(H_c\) the positive root of \({{\mathcal {Z}}}'\) , that is,

On \([H_c,+\infty )\) , \({{\mathcal {Z}}}\) is increasing.

We directly borrow from [ 37 , Section 4.1] that when \(F>(\nu +1)/\nu ^2\) , one has \(H_*>H_c\) and, when moreover \(H_*\le H_L\) , \({{\mathcal {Z}}}(H_*)>0\) . This directly implies that in cases (iii) and (iv) of Proposition  2.1 , indeed \(\inf _{{\mathbb {R}}_-^*}\left( -f_4+\frac{1}{2}f_1f_2+\frac{1}{2}f_1'\right) >0\) .

To complete the analysis of non real eigenvalues, we only need to show that in case (ii), \(H_L>H_c\) and \({{\mathcal {Z}}}(H_L)>0\) . It is straightforward to check that when \(F\le \sqrt{2\nu (\nu +1)}\) indeed \(H_L>H_c\) , whereas \({{\mathcal {Z}}}(H_L)>0\) is exactly equivalent to \(F< \sqrt{2\nu (\nu +1)}\) .

This achieves the proof that a spectral gap is present for non real eigenvalues.

4.3 Real Growth Rates

We now turn our attention to the case of real eigenvalues. Throughout the present subsection, we assume that \(\lambda \in {\mathbb {R}}_+\) and our goal is again to rule out the possibility of a nonzero w in \(H^2({\mathbb {R}}_-^*;{\mathbb {C}})\) solving ( 4.1 )–( 4.2 ).

Our starting point is again Equation ( 4.3 ), that we write now as

where \({{\mathcal {A}}}_\lambda \) and \({{\mathcal {B}}}\) are the symmetric sesquilinear Footnote 7 forms on \(H^1({\mathbb {R}}_-^*;{\mathbb {C}})\) defined through their quadratic forms

Note that since \(\lambda \in {\mathbb {R}}_+\) , the analysis of the former subsection yields that \({{\mathcal {A}}}_\lambda \) is positive definite when \(F< \sqrt{2\nu (\nu +1)}\) . In order to conclude it is therefore sufficient to prove that \({{\mathcal {B}}}\) is also positive definite, that is,

The equivalence between the two conditions follows from the following Gårding-type inequality: there exist positive c and C such that, for any \(v\in H^1\) , \({{\mathcal {B}}}(v,v)\ge c\Vert v\Vert _{H^1}^2-C\Vert v\Vert _{L^2}^2\) . A refined version of the latter is proved below.

As in [ 37 ], we prove the latter by a continuity/homotopy argument. To set it, we introduce, for \(x_0\in {\mathbb {R}}_-\) , \({{\mathcal {B}}}_{x_0}\) the symmetric sesquilinear form on \(H^1({\mathbb {R}}_-^*;{\mathbb {C}})\) defined through its quadratic form

The explicit definition of \(c_2^{[x_0]}\) is given below but let us already anticipate that our choice ensures that \(c_2^{[x_0]}\) depends smoothly on \(x_0\) and converges as \(x_0\rightarrow -\infty \) to a negative value. Note moreover that

This implies that when \(x_0\) is sufficiently close to \(-\infty \) , \({{\mathcal {B}}}_{x_0}\) is positive definite. To motivate the expression for \(c_2^{[x_0]}\) , we first observe that

then, consistently we set

Note that as announced

The continuity argument is applied to the continuous function

The fact that the foregoing function is indeed defined follows again from the Gårding inequality mentioned above. To complete our study of cases (ii) and (iv), it is sufficient to prove that this function cannot vanish. This follows in a straightforward way from the series of two lemmas stated and proved below.

To prepare the lemmas, we first quantify the possible failure of coercivity. For any

there exist positive \(\eta _\kappa \) , \(c_\kappa \) and \(C_\kappa \) such that for any \(x_0\in {\mathbb {R}}_-\) and \(v\in H^1\) ,

Indeed \(\eta _\kappa \) may be chosen by imposing

and the existence of \(c_\kappa \) and \(C_\kappa \) is a consequence of rough bounds on coefficients and the following Sobolev inequality,

that holds for any \(\eta >0\) (with an implicit constant depending on \(\eta \) ).

As a second and last preliminary to lemmas, we find it convenient to explicitly introduce the self-adjoint operator on \(L^2({\mathbb {R}}_-^*;{\mathbb {C}})\) , \({{\mathcal {L}}}_{x_0}\) , of domain denoted \(D_{x_0}\) , associated with \({{\mathcal {B}}}_{x_0}\) . Explicitly

and for \(v\in D_{x_0}\) ,

then there exists \(v\in D_{x_0}\) , \(v\not \equiv 0\) , such that \({{\mathcal {L}}}_{x_0}v=0\) .

Let us consider \((v_k)_{k\in {{\mathbb {N}}}}\) a minimizing sequence, normalized by \(\Vert v_k\Vert _{L^2}=1\) . From the Gårding estimate, we know that \((v_k)_{k\in {{\mathbb {N}}}}\) is bounded in \(H^1\) and thus, up to extracting a subsequence, we may assume that \((v_k)_{k\in {{\mathbb {N}}}}\) converges weakly in \(H^1\) to some \(v_\infty \in H^1\) . As a direct consequence of the Hahn-Banach theorem, we deduce that

Now pick \(\eta >0\) such that \(f_2^2+2f_2'\) is positive outside \((-\eta ,0)\) and note that since \(H^1((-\eta ,0))\) is compactly embedded in \(L^2((-\eta ,0))\) we may assume that \((v_k)_k\) converges strongly to \(v_{\infty }\) in \(L^2((-\eta ,0))\) . Combined with ( 4.4 ), this is sufficient to take the limit \(k\rightarrow \infty \) in the remaining part of \({{\mathcal {B}}}_{x_0}(v_k,v_k)\) . As a result

We now prove that \(v_\infty \) is nonzero. This is the place where we use the refined version of the Gårding estimate. Indeed it implies that there exist positive \(\eta '\) and K such that when k is sufficiently large so as to force that \({{\mathcal {B}}}_{x_0}(v_k,v_k)\) is sufficiently small

Since \(\Vert v_k\Vert _{L^2}=1\) , we deduce that

Then, since we may assume that \((v_k)_k\) converges strongly to \(v_{\infty }\) in \(L^2((-\eta ',0))\) , we conclude that \(v_\infty \) is nonzero.

Let us set \(v:=v_\infty /\Vert v_\infty \Vert _{L^2}\) . The vector v is nonzero and satisfies \({{\mathcal {B}}}_{x_0}(v,v)\le 0\) , thus \({{\mathcal {B}}}_{x_0}(v,v)=0\) . Since v minimizes the quadratic form associated with \({{\mathcal {B}}}_{x_0}\) among vectors of \(H^1\) with unit \(L^2\) norm, there exists \(\mu \in {\mathbb {C}}\) such that \({{\mathcal {B}}}_{x_0}(v,\cdot )\,=\,\mu \,\langle v,\cdot \rangle _{L^2}\) . In particular \(v\in D_{x_0}\) and \({{\mathcal {L}}}_{x_0}v=\overline{\mu } v\) . Since v is nonzero, evaluating the relation at v shows that \(\mu =0\) and concludes the proof of the lemma. \(\square \)

The foregoing lemma is very close to many standard results but, unfortunately, we have not found a directly applicable version in the literature. Hence the above proof.

If \(v\in D_{x_0}\) is such that \({{\mathcal {L}}}_{x_0}v=0\) then \(v\equiv 0\) .

Note that the set of \(v\in H^2({\mathbb {R}}_-)\) such that

is one-dimensional. Moreover from the fact that \((H',Q')\) solves the interior spectral ODE system in original formulation, we deduce that \((H-H_L)''=-f_2\,(H-H_L)'\) , and thus that

spans the above set.

To conclude we just need to check that \(v^{[x_0]}\notin D_{x_0}\) . This is indeed the case since

\(\square \)

4.4 The Riemann Shock Case

We conclude our stability analysis by discussing how to adapt the above arguments to the Riemann shock case. The overall strategy is identical but details should be changed at various places.

We only indicate these modifications. To begin with, since \([r(W)]=0\) , it is convenient to replace \(\psi \) with \({\widetilde{\psi }}:=\lambda \psi \) . This does not change the nature of the spectral problem when \(\lambda \ne 0\) and simply decreases by 1 the algebraic multiplicity of the eigenvalue \(\lambda =0\) . Our task is thus to determine when there exists a nonzero \((v,{\widetilde{\psi }})\in H^1_{\eta _L}({\mathbb {R}}_-^*;{\mathbb {C}}^2)\times {\mathbb {C}}\) solving

Since [ W ] is non zero, one may eliminate \({\widetilde{\psi }}\) and reduce the discussion to the existence of a nonzero \(u=(u_1,u_2):=(v_1,-cv_1+v_2)\) in \(H^1_{\eta _L}({\mathbb {R}}_-^*;{\mathbb {C}}^2)\) such that on \({\mathbb {R}}_-^*\)

There is no nonzero solution with either \(u_2\) vanishing identically or \(\lambda =0\) , so that the problem is equivalent to finding a nonzero \(u_2\in H^2_{\eta _L}({\mathbb {R}}_-^*;{\mathbb {C}})\) solving on \({\mathbb {R}}_-^*\)

with H and Q constant equal to \(H_L\) and \(H_L^{3/2}\) respectively.

From here no change is needed in the reduction from \(u_2\) to w , nor in Subsection  4.2 . The core of Subsection  4.3 is simply replaced with a direct check that \({{\mathcal {B}}}\) is positive definite. This follows from the fact that \(f_2\) is a negatively-valued constant function and \(c_2\) is also negative since it is equal to one half of this value. The sign observation stems from \(H_L>H_s\) and \(F>2\) which imply

Summarizing the results of the present section with the ones of Sect.  3 , we obtain the following proposition:

Proposition 4.3

5 linear and nonlinear convective stability.

At this point, we have shown that convective spectral stability holds (with scalar weight) for \(F<\sqrt{2\nu (\nu +1)}\) , and fails (for any weight) for \(F>\sqrt{2\nu (\nu +1)}\) , We now complete our discussion of convective stability by invoking a Lyapunov-type argument showing that convective spectral stability implies linear and nonlinear convective orbital stability, at time-exponential rate.

Convective spectral stability in the semilinear parabolic case, with a smooth background traveling wave, yields fairly immediately time-exponential asymptotic orbital stability, by well known arguments of Sattinger [ 35 ] and Henry [ 19 ] similar to those for the finite-dimensional ODE case. The present setting involving discontinuous background waves and quasilinear hyperbolic equations requires a much more technical analysis, at the frontier of the current knowledge on nonlinear wave stability theory.

The expository choice we make is to borrow results from the forthcoming [ 15 ] that carries out a systematic development in a more general setting, in the spirit of [ 14 ], where Propositions  5.1 and  5.2 below are proved. We stress however that, to a large extent, a relatively simple adaptation of the techniques used in [ 39 ] for the neutrally stable case. Footnote 8 would already be sufficient for the present case. Nevertheless a self-contained exposition of this adequate version would essentially double the size of the present contribution. Even for the smooth case, none of the results in the literature seems directly applicable to yield the time-exponential stability proved below, but, likewise, a relatively simple variation on the neutrally-stable treatments [ 26 , 27 ] would yield the required result.

5.1 Linear Estimates

We begin with estimates for the linearized problem ( 3.3 ),

with initial data \(v_0(x)\) and interior source term F ( t ,  x ).

Proposition 5.1

Let \(W=(H,Q)\) be a traveling-wave solution of type (v) and consider \((\eta _L,\eta _R)\) spatial weight growths ensuring Footnote 9 a spectral gap. Then there exist positive \(\theta \) and C such that for any \(1\le p\le \infty \) , if

then for v the unique mild solution to ( 5.1 ) (in \({{\mathcal {C}}}^0({\mathbb {R}}_+;L^p_{\eta _L,\eta _R}({\mathbb {R}}))\) if \(1\le p<\infty \) , \({{\mathcal {C}}}^0({\mathbb {R}}_+;BUC^0_{\eta _L,\eta _R}({\mathbb {R}}))\) if \(p=\infty \) ) with initial data \(v_0\) , there exists a phase shift \(\varphi \in {{\mathcal {C}}}^1({\mathbb {R}}_+)\) vanishing initially such that for any \(t\ge 0\)

In the foregoing statement \(BUC^0(\Omega )\) denotes the space of functions that are bounded on \(\Omega \) , and uniformly continuous on each connected component of \(\Omega \) .

We recall that Duhamel formula enables one to reduce the previous statement to the sourceless case. Moreover we point out that in the case \(p=2\) the statement follows from the Gearhart–Prüss theorem and high-frequency bounds on resolvents.

Consider again the linearized problem ( 3.4 ):

with initial data \((v_0(x),\psi _0)\) , interior source term F ( t ,  x ), and boundary source-term G ( t ). Recall from the original derivation of ( 3.4 ) that here there is no freedom in the phase shift that may be removed from v so as to obtain time decay. We need to prove that \(v-(-\psi )\,W'\) is decaying. In contrast, in the smooth case, the phase shift \(\varphi \) is far from unique.

Proposition 5.2

Let \(W=(H,Q)\) be a traveling-wave solution of type (ii)-(iv) satisfying the sharp convective spectral stability condition Footnote 10

and consider \((\eta _L,\eta _R)\) spatial weight growths ensuring, Footnote 11 a spectral gap. Then there exist positive \(\theta \) and C such that for any \(1\le p\le \infty \) , if

then \((v,\psi )\) , the unique mild solution to ( 5.3 ) with initial data \((v_0,\psi _0)\) , satisfies for any \(t\ge 0\)

Note that the level of regularity of the previous statement is insufficient, alone, to define traces at \(0^\pm \) . The existence of those is a consequence of the fact that \((v,\psi )\) solves ( 5.3 ) and that the shock is non characteristic.

5.2 Nonlinear Stability

Using Propositions  5.1 and  5.2 in order to prove nonlinear stability results induces a severe loss of derivatives due to the quasilinear character of the original system. A by-now classical way to cure this loss is to combine the latter with nonlinear high-frequency damping estimates, that show that as long as the Lipschitz norm of the solution remains under control, the time decay of any Sobolev norm is slaved to the time decay of the \(L^2\) norm. Designing such nonlinear high-frequency damping estimates is a significant part of the nonlinear stability analysis. When proceeding in this way, it is actually sufficient to prove linear stability with derivative losses, as accessible through what the fourth author has dubbed the “poor man’s Prüss construction” [ 42 ]. On nonlinear high-frequency damping estimates, we refer to [ 29 , Appendix A] for an introduction the classical Kawashima version for the stability of constant states [ 23 , 36 ] and to [ 14 , 27 , 32 , 39 ] for versions more directly related to the present analysis.

Theorem 5.3

Let \(W=(H,Q)\) be a traveling-wave solution of type (v) and consider \((\eta _L,\eta _R)\) spatial weight growths ensuring Footnote 12 a spectral gap, with \(\eta _L\) positive and \(\eta _R\) negative. Then there exist positive \(\delta \) , \(\theta \) and C such that if

then for w the unique mild solution to ( 1.1 ) (in \({{\mathcal {C}}}^0({\mathbb {R}}_+;H^2_{\eta _L,\eta _R}({\mathbb {R}}))\) ) with initial data \(w_0\) , there exists a phase shift \(\varphi \in {{\mathcal {C}}}^1({\mathbb {R}}_+)\) vanishing initially and an asymptotic shift \(\varphi _\infty \in {\mathbb {R}}\) such that for any \(t\ge 0\)

and \(|\varphi _\infty |\le C\,\delta _{w_0}\) .

Theorem 5.4

Let \(W=(H,Q)\) be a traveling-wave solution of type (ii)-(iv) satisfying the sharp convective spectral stability condition Footnote 13

and consider \((\eta _L,\eta _R)\) spatial weight growths ensuring, Footnote 14 a spectral gap, with \(\eta _L\) positive and \(\eta _R\) negative. Then there exist positive \(\delta \) , \(\theta \) and C such that if

with \(w_0-W\) supported away from zero, then there exists a global solution to ( 1.1 ), w , with initial data \(w_0\) having at each time \(t\ge 0\) a single shock, located at \(ct+\psi (t)\) , with \(\psi \in {{\mathcal {C}}}^1({\mathbb {R}}_+)\) vanishing initially and an asymptotic shift \(\psi _\infty \in {\mathbb {R}}\) such that for any \(t\ge 0\)

and \(|\psi _\infty |\le C\,\delta _{w_0}\) .

Note that none of the constants depend on how far the support of \(w_0-W\) is from 0. The assumption is simply made to assure that the initial data is compatible with the short-time persistence of a single-shock piecewise- \(H^2\) solution. We could have instead assumed directly the optimal but cumbersome compatibility conditions, as in [ 14 , 15 ]. On the related local-in-time propagation of regularity we refer to [ 6 , 28 ].

6 Numerical Time-Evolution Experiments

figure 5

Left panel: Dambreak initial data ( 6.1 ); Right panel: Dambreak initial data ( 6.2 )

figure 6

Space-time plots of simulations shown in Figs.  7 , 8 , 9 and 10

6.1 Nonmonotone Discontinuous Waves and Riemann Shock

Throughout this section we set \(H_L=1\) and \(H_R=0.7\) . We first note that profiles are of nonmonotone discontinuous type (ii) if

and they are convectively exponentially stable given that

When \(F=\frac{\sqrt{85}}{10}+\frac{\sqrt{238}}{14}\) profiles are of Riemann shock type (iii) and convectively exponentially stable. In all cases, speed of the waves is given by

For our numerical experiments, we use a perturbed dambreak initial data given by

See Fig.  5 left panel for a plot of ( 6.1 ).

figure 7

Numerical simulation of ( 1.1 ) with \(F=2.28\) and initial data ( 6.1 )

Note that the slower speed of the invading front can be heuristically predicted by tracking how the absolute spectrum associated with \(H_L\) depends on the speed of the moving frame in which it is computed. To be more concrete, we revisit computations from Sect.  3.2 by allowing the speed of the moving frame s to vary (instead of being fixed to c , the wave speed of the traveling wave of interest) and correspondingly mark with s different quantities introduced there. We are interested in the absolute spectrum of \(H_L\) at speed s , that is in the \(\lambda \) such that the real parts of the spatial eigenvalues of \(G_{H_L,s}(\lambda )\) coincide, \(\textrm{Re}(\gamma _{-,H_L,s}(\lambda ))=\textrm{Re}(\gamma _{+,H_L,s}(\lambda ))\) . Following [ 13 ], we may define a so-called absolute spreading speed \(s_\textrm{abs}\) as the infimum of wave speeds s for which the absolute spectrum remains unstable in the moving frame s . Computations from Sect.  3.2 yield

In our case, with \(F=2.30\) , \(H_L=1\) and \(H_R=0.7\) , we have \(s_\textrm{abs}-c=-0.00305\ldots \) . The dashed line in the second panel of Fig.  6 has precisely a slope given by \(s_\textrm{abs}-c\) which matches the onset of the invading roll waves quite accurately.

figure 8

Numerical simulation of ( 1.1 ) with \(F=2.30\) and initial data ( 6.1 )

figure 9

Numerical simulation of ( 1.1 ) with \(F=\frac{\sqrt{85}}{10}+\frac{\sqrt{238}}{14}\) and initial data ( 6.1 )

6.2 Increasing Smooth Waves

Finally, we make \(H_R>H_L\) to test if the corresponding increasing smooth “reverse shock” can be the large-time asymptotic limit. We fix \(H_L=1\) and \(H_R=1.3\) such that profiles are of increasing smooth type (i) if

See Fig.  5 right panel for a plot of ( 6.2 ). We report an “invading back” connecting roll wave patterns on the left to a constant state on the right. See Fig.  10 for plots at time 5, 25, and 100 of fluid heights in the comoving frame \(c=\frac{H_L+\sqrt{H_LH_R}+H_R}{\sqrt{H_L}+\sqrt{H_R}}\sim 1.6074\) and Fig.  6 last panel for a space-time plot. This illustrates once again that our (local) stability conditions indeed successfully predict large-scale asymptotic behavior.

We also tested the predictive feature of the absolute spreading speed introduced in the convectively unstable case beyond its expected range of validity by computing \(s_\textrm{abs}\) in the present case and found \(s_\textrm{abs}-c=-0.3852\ldots \) . Quite surprisingly and remarkably, this predicted speed compares well with the speed of the primary invading front (see the dashed line in the forth panel of Fig.  6 ) for short time.

figure 10

Numerical simulation of ( 1.1 ) with \(F=3\) and initial data ( 6.2 )

Roll Wave Selection . So far we are lacking even a heuristic argument to predict which roll wave is selected in the invading front pattern. Let us recall that, even when the translational invariance is quotiented, roll waves form a two-parameter family.

By integrating over a large space-time box the conservation law of (SV), one may derive formally a constraint equation

where \(\langle A\rangle \) denotes the average of the quantity A over one period of the roll wave pattern. Yet this is one equation short to fully identify the roll pattern, leaving a degree of freedom still to be determined.

Looking toward the future, we would like to add two more comments on this question. First, we point out that we have estimated numerically the wave period and wave speed of the observed roll pattern (as approximately 1 and 1.4 respectively) and checked that the wave does lie in the stability region of the roll-wave stability diagram [ 21 , Fig. 3(c)].

Second, we mention that the oscillatory instability pattern between \(H_L\) and the roll-wave seems to be expanding linearly in time, preventing a direct connection from the roll-wave to \(H_L\) . One can not exclude that the identification of the missing roll parameter requires a deep understanding of this pattern in a way reminiscent of the resolution of the Gurevich–Pitaevskii problem through the analysis of dispersive shocks [ 5 ].

A complete, rigorous treatment of this bifurcation would be very interesting to carry out.

Data availability

Data will be made available on reasonable request.

It was incorrectly stated there, as a side remark, that for hydrodynamically unstable endstates, the only hydraulic shock profiles were smooth, “reverse”-direction shocks not connected to equilibrium dynamics.

In both directions.

The first part corresponds to the Riemann shock case.

For the moment we have only proved stabilization of the essential spectrum but we do prove full stability in the end.

With obvious notational adaptation for weighted spaces.

It should not be confused with the w used in the initial introduction of the spectral problem.

Consistently with our convention for scalar products, they are linear in their second factors.

See also the related discussion of [ 39 , p. 201].

For instance, \(\eta _L\) positive and \(\eta _R\) negative both sufficiently small in absolute value.

Automatically satisfied in cases (iii) and (iv).

For instance, when \(F>2\) , \(\eta _L\in (\eta _L^\textrm{min}\eta _L^\textrm{max})\) and \(\eta _R\) sufficiently negative; when \(F<2\) , \(\eta _L\) positive and \(\eta _R\) negative both sufficiently small in absolute value.

Arendt, W. , Batty, C.J.K. , Hieber, M. , Neubrander, F. : Vector-Valued Laplace Transforms and Cauchy Problems, Volume 96 of Monographs in Mathematics, 2nd edn. Birkhäuser, Basel (2011)

Book   Google Scholar  

Barker, B. , Johnson, M.A. , Noble, P. , Rodrigues, L.M. , Zumbrun, K. : Note on the stability of viscous roll-waves. Comptes Rendus Mécanique 345 (2), 125–129, 2017

Article   ADS   Google Scholar  

Barker, B. , Johnson, M.A. , Noble, P. , Rodrigues, L.M. , Zumbrun, K. : Stability of viscous St.Venant roll waves: from onset to infinite Froude number limit. J. Nonlinear Sci. 27 (1), 285–342, 2017

Article   ADS   MathSciNet   Google Scholar  

Barker, B. , Johnson, M.A. , Rodrigues, L.M. , Zumbrun, K. : Metastability of solitary roll wave solutions of the St.Venant equations with viscosity. Phys. D 240 (16), 1289–1310, 2011

Article   MathSciNet   Google Scholar  

Benzoni-Gavage, S. , Mietka, C. , Rodrigues, L.M. : Modulated equations of Hamiltonian PDEs and dispersive shocks. Nonlinearity 34 (1), 578–641, 2021

Benzoni-Gavage, S. , Serre, D. : Multidimensional Hyperbolic Partial Differential Equations. Oxford Mathematical Monographs. The Clarendon Press, Oxford (2007)

Google Scholar  

Blochas, P. , Rodrigues, L.M. : Uniform asymptotic stability for convection–reaction–diffusion equations in the inviscid limit towards Riemann shocks. Ann. Inst. H. Poincaré C Anal. Non Linéaire (to appear)

Bressan, A. : Hyperbolic Systems of Conservation Laws, Volume 20 of Oxford Lecture Series in Mathematics and its Applications. Oxford University Press, Oxford (2000)

Clawpack Development Team . Clawpack version 5.4.0 (2017)

Dressler, R.F. : Mathematical solution of the problem of roll-waves in inclined open channels. Commun. Pure Appl. Math. 2 , 149–194, 1949

Duchêne, V. , Rodrigues, L.M. : Large-time asymptotic stability of Riemann shocks of scalar balance laws. SIAM J. Math. Anal. 52 (1), 792–820, 2020

Duchêne, V. , Rodrigues, L.M. : Stability and instability in scalar balance laws: fronts and periodic waves. Anal. PDE 15 (7), 1807–1859, 2022

Faye, G. , Holzer, M. , Scheel, A. , Siemer, L. : Invasion into remnant instability: a case study of front dynamics. Indiana Univ. Math. J. 71 (5), 1819–1896, 2022

Faye, G. , Rodrigues, L.M. : Exponential asymptotic stability of Riemann shocks of hyperbolic systems of balance laws. SIAM J. Math. Anal. (to appear)

Faye, G. , Rodrigues, L. M. : Convective stability and exponential asymptotic stability of non constant shocks of hyperbolic systems of balance laws (work in progress)

Garénaux, L. , Rodrigues,L. M. : Convective stability in scalar balance laws (work in progress)

Godillon, P. : Linear stability of shock profiles for systems of conservation laws with semi-linear relaxation. Phys. D 148 (3–4), 289–316, 2001

Godillon, P. , Lorin, E. : A Lax shock profile satisfying a sufficient condition of spectral instability. J. Math. Anal. Appl. 283 (1), 12–24, 2003

Henry, D. : Geometric theory of Semilinear Parabolic Equations, volume 840 of Lecture Notes in Mathematics. Springer, Berlin (1981)

Jeffreys, H. : The flow of water in an inclined channel of rectangular section. Philos. Mag. 49 (293), 793–807, 1925

Article   Google Scholar  

Johnson, M.A. , Noble, P. , Rodrigues, L.M. , Yang, Z. , Zumbrun, K. : Spectral stability of inviscid roll waves. Commun. Math. Phys. 367 (1), 265–316, 2019

Kapitula, T. , Promislow, K. : Spectral and Dynamical Stability of Nonlinear Waves, Volume 185 of Applied Mathematical Sciences. Springer, New York (2013)

Kawashima, S. : Systems of a hyperbolic–parabolic composite type, with applications to the equations of magnetohydrodynamics. Ph.D. thesis, Kyoto University (1983)

Liu, T.-P. : Hyperbolic conservation laws with relaxation. Commun. Math. Phys. 108 (1), 153–175, 1987

Mandli, K. , Ahmadia, A. , Berger, M. , Calhoun, D. , George, D. , Hadjimichael, Y. , Ketcheson, D. , Lemoine, G. , LeVeque, R. : Clawpack: building an open source ecosystem for solving hyperbolic pdes. PeerJ Comput. Sci. 2 , e68, 2016

Mascia, C. , Zumbrun, K. : Pointwise Green’s function bounds and stability of relaxation shocks. Indiana Univ. Math. J. 51 (4), 773–904, 2002

Mascia, C. , Zumbrun, K. : Stability of large-amplitude shock profiles of general relaxation systems. SIAM J. Math. Anal. 37 (3), 889–913, 2005

Métivier, G. : Stability of multidimensional shocks. In Advances in the Theory of Shock Waves, Volume 47 of Programming Nonlinear Differential Equations Application, pages 25–103. Birkhäuser Boston, Boston, MA (2001)

Rodrigues, L.M. : Asymptotic stability and modulation of periodic wavetrains. General theory & applications to thin film flows. Habilitation à diriger des recherches, Université Lyon 1, 2013

Rodrigues, L.M. , Yang, Z. , Zumbrun, K. : Convective-wave solutions of the Richard–Gavrilyuk model for inclined shallow-water flow. Water Waves 5 (1), 1–39, 2023

Rodrigues, L.M. , Yang, Z. , Zumbrun, K. : Existence and stability of hydraulic shock profiles for the Richard–Gavrilyuk model (work in progress)

Rodrigues, L.M. , Zumbrun, K. : Periodic-coefficient damping estimates, and stability of large-amplitude roll waves in inclined thin film flow. SIAM J. Math. Anal. 48 (1), 268–280, 2016

Sandstede, B. : Stability of travelling waves. In Handbook of dynamical systems, Vol. 2 , pages 983–1055. North-Holland, Amsterdam (2002)

Sandstede, B. , Scheel, A. : Essential instabilities of fronts: bifurcation, and bifurcation failure. Dyn. Syst. 16 , 1–28, 2001

Sattinger, D. : On the stability of waves of nonlinear parabolic systems. Adv. Math. 22 (3), 312–355, 1976

Shizuta, Y. , Kawashima, S. : Systems of equations of hyperbolic-parabolic type with applications to the discrete Boltzmann equation. Hokkaido Math. J. 14 (2), 249–275, 1985

Sukhtayev, A. , Yang, Z. , Zumbrun, K. : Spectral stability of hydraulic shock profiles. Phys. D 405 , 132360, 2020

Texier, B. , Zumbrun, K. : Entropy criteria and stability of extreme shocks: a remark on a paper of Leger and Vasseur. Proc. Am. Math. Soc. 143 (2), 749–754, 2015

Yang, Z. , Zumbrun, K. : Stability of hydraulic shock profiles. Arch. Ration. Mech. Anal. 235 (1), 195–285, 2020

Zumbrun, K. : Multidimensional stability of planar viscous shock waves. In Advances in the Theory of Shock Waves, Volume 47 of Programming Nonlinear Differential Equations Application, pages 307–516. Birkhäuser, Boston, MA (2001)

Zumbrun, K. , Howard, P. : Pointwise semigroup methods and stability of viscous shock waves. Indiana Univ. Math. J. 47 (3), 741–871, 1998

Zumbrun, K. , Jenssen, H.K. , Lyng, G. : Chapter 5 - stability of large-amplitude shock waves. Volume 3 of Compressible Navier–Stokes Equations Handbook of Mathematical Fluid Dynamics. North-Holland, Amsterdam (2005)

Download references

Author information

Authors and affiliations.

Institut de Mathématiques de Toulouse, UMR5219, CNRS UPS IMT, Université de Toulouse, 31062, Toulouse Cedex 9, France

Grégory Faye

CNRS, IRMAR - UMR 6625, Univ Rennes & IUF, 35000, Rennes, France

L. Miguel Rodrigues

Academy of Mathematics and Systems Science, Chinese Academy of Sciences, Beijing, 100190, China

Indiana University, Bloomington, IN, 47405, USA

Kevin Zumbrun

You can also search for this author in PubMed   Google Scholar

Corresponding author

Correspondence to Grégory Faye .

Additional information

Communicated by J. Bedrossian.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

G.F. acknowledges support from the ANR via the project Indyana under Grant Agreement ANR-21-CE40-0008 and Labex CIMI under Grant Agreement ANR-11-LABX-0040. Research of L.M.R. was partially supported by EPSRC Grant No. EP/R014604/1. Research of K.Z. was partially supported under NSF Grant Nos. DMS-2154387 and DMS-220610.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/ .

Reprints and permissions

About this article

Faye, G., Rodrigues, L.M., Yang, Z. et al. Existence and Stability of Nonmonotone Hydraulic Shocks for the Saint Venant Equations of Inclined Thin-Film Flow. Arch Rational Mech Anal 248 , 82 (2024). https://doi.org/10.1007/s00205-024-02033-4

Download citation

Received : 19 July 2023

Accepted : 23 August 2024

Published : 13 September 2024

DOI : https://doi.org/10.1007/s00205-024-02033-4

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

  • Find a journal
  • Publish with us
  • Track your research

COMMENTS

  1. Froude number

    In continuum mechanics, the Froude number (Fr, after William Froude, / ˈ f r uː d / [1]) is a dimensionless number defined as the ratio of the flow inertia to the external force field (the latter in many applications simply due to gravity).The Froude number is based on the speed-length ratio which he defined as: [2] [3] = where u is the local flow velocity (in m/s), g is the local gravity ...

  2. Froude Number

    Froude's experiments were validated in full-scale trials, and his methods of ship scale modeling are still used today. Froude's work on free-surface flows was acknowledged by naming the ratio of inertial and gravitational forces as the "Froude number." It is one of two numbers in fluid mechanics that dominate the scaling of all flows.

  3. William Froude

    William Froude (/ ˈ f r uː d /; [1] 28 November 1810 in Devon [2] - 4 May 1879 in Simonstown, Cape Colony) was an English engineer, hydrodynamicist and naval architect. He was the first to formulate reliable laws for the resistance that water offers to ships (such as the hull speed equation) and for predicting their stability.

  4. Professor Robert B. Laughlin, Department of Physics, Stanford University

    The Froude number is the ratio of flow velocity to wave celerity and marks the boundary between critical and subcritical flow [1]. For conditions in which the Froude number is greater than one, flow velocity exceedes the wave celerity (the speed of an individual wave crest) and the fluid motion is smooth. ... Contemporary experiments have put ...

  5. Fluid flow: Froude and Reynolds numbers

    Froude number. Froude's influential paper of 1861 was published by the Institute of Naval Architects (PDF available). Froude had surmised that, to predict the behaviour of a ship moving through water, he would need to experiment with much smaller versions of ships, or models, that could be scaled to the behaviour of much larger vessels. Thus ...

  6. William Froude: the father of hydrodynamics

    Although Froude died in 1879 at the age of 68, his son Robert Edmund (Eddie) Froude continued the research and established a new experimental facility on a redundant plot adjacent to the naval gunboat yard at Haslar. The last test at Torquay was completed on 5th January 1886 and the new tank at the Admiralty Experiment Works, Haslar opened a ...

  7. Froude number (Fr)

    Froude number (Fr), in hydrology and fluid mechanics, dimensionless quantity used to indicate the influence of gravity on fluid motion. It is generally expressed as Fr = v /( gd ) 1 / 2 , in which d is depth of flow, g is the gravitational acceleration (equal to the specific weight of the water divided by its density , in fluid mechanics), v is ...

  8. PDF Chapter 2 Flow Past a Sphere I: Dimensional Analysis, Reynolds Numbers

    REYNOLDS NUMBERS, AND FROUDE NUMBERS INTRODUCTION 1 Steady flow past a solid sphere is important in many situations, both in the natural environment and in the world of technology, and it serves as a good ... obtain the function by experiment, as is commonly the case in problems of flow of real fluids. With flow past the sphere as an example we ...

  9. William Froude

    The towing carriage met Froude's desire for something more scientific for his new experiment tank, something that would pull the models at a steady and measurable rate. It utilized a 3-foot, 3-inch gauge railway for the towing equipment and screw dynamometer, "carried by the roof principals and extended the full length of the waterway.

  10. Froude number

    The Froude number, an important parameter for the study of liquids moving in a free surface (e.g., surface wave motion), is defined as F r = v 2 / gL or F r = ρv 2 L 2 /ρgL 3.Because ρL 3 = m (mass), one can also have F r = ρv 2 L 2 /mg.Since ρv 2 is the inertial force per unit area, the numerator of this last formula represents the total inertial force of the fluid.

  11. Froude Similarity

    Froude similarity considers besides inertia the gravity force, which is dominant in most free surface flows, especially if friction effects are negligible or for highly turbulent phenomena such as wave breaking. The Froude similarity requires identical Froude numbers between model and its prototype for each selected experiment.

  12. Who is faster, the currents or the waves? The Froude number

    For an example of the explanatory power of the Froude number, you see a tank experiment we did a couple of years ago with Rolf Käse and Martin Vogt . There is actually a little too much going on in that tank for our purposes right now, but the ridge on the right can be interpreted as, for example, the Greenland-Scotland-Ridge, making the blue ...

  13. Froude Number and Flow States

    Froude Number and Flow States. The Froude number, Fr, is a dimensionless value that describes different flow regimes of open channel flow. The Froude number is a ratio of inertial and gravitational forces. · Gravity (numerator) - moves water downhill. · Inertia (denominator) - reflects its willingness to do so. Where: V = Water velocity.

  14. Froude number

    The Froude number is a dimensionless parameter used to determine the similarity between the flow of fluids at different scales. It is named after William Froude, a British engineer who studied the movement of ships and waves in water. The Froude number is defined as the ratio of the inertia forces to the gravitational forces acting on a fluid ...

  15. 2.7: Significance of Reynolds Numbers and Froude Numbers

    The first term in Equation 2.7.2 2.7.2 is then proportional to μVL/ρL2V2 μ V L / ρ L 2 V 2, or μ/ρLV μ / ρ L V. This is simply the inverse of a Reynolds number. The Reynolds number in any fluid problem is therefore inversely proportional to the ratio of a viscous force and a quantity with the dimensions of a force, the rate of change of ...

  16. PDF General Modelling and Scaling Laws

    Froude number . Fn • The ratio between inertia and gravity: gL • Dynamic similarity requirement between model and full scale: • Equality in . Fn. in model and full scale will ensure that gravity forces are correctly scaled • Surface waves are gravity -driven ⇒ equality in . Fn. will ensure that wave resistance and other wave forces are

  17. Froude number

    Froude number. One of the criteria for similarity of the motion of a liquid or gas, applicable in cases when the influence of gravitational forces is significant. The Froude number characterizes the relationship between the inertial and gravitational forces acting on an elementary volume of the liquid or gas. The Froude number is.

  18. William Froude

    Froude's writings include "Experiments on the Surface-friction Experienced by a Plane Moving Through Water," in British Association for the Advancement of Science Report, 42nd Meeting, 1872; and "On Experiments with H.M.S. Greyhound," in Transactions of the Institution of Naval Architects, 16 (1874), 36-73.

  19. Froude Number

    The Froude Number is often depth-averaged or applied as a densiometric function to compare experiments of sediment gravity flows in laboratory tanks, such as Euro Tank (sensu Pohl, 2019). Lower stage plane beds and ripples are common bedforms in fine-grained deepwater environments that are formed by subcritical flows, developed beneath a near ...

  20. Froude Number: Equation & Subcritical Flow

    C. The Froude Number equation signifies the importance of gravitational forces in fluid mechanics. The variables are mass of the fluid, acceleration due to gravity and the depth of flow. D. The Froude Number equation is based on Buckingham's Pi theory and it represents the force applied on a fluid under various conditions of flow rates and ...

  21. Hydraulic Jump through a Sluice Gate

    A Froude number of 2.5 to 4.5 is the worst design range, as the jump in this range will create large waves that could cause structural damage. Theory: The performance of a hydraulic jump depends mainly on the value of the Froude number. The Froude number is defined as ... In this experiment, the velocity is varied by varying the speed of the ...

  22. PDF MASSACHUSETTS INSTITUTE OF TECHNOLOGY Department of Mechanical

    Froude approach. In 1978 the International Towing Tank Congress (ITTC) adopted a updated standardized procedure for performing ship resistance experiments and calculations. The ITTC-78 method considers the total model resistance as the sum of a viscous component and a wave making component (see reference material).

  23. Froude Number

    It is expressed as, Froude No = v √gd. Here, 'd' represents the depth of flow, 'g' is the acceleration due to gravity, 'v' is the celerity or gravity of the small surface wave. When Froude No. is at a critical point that is Fr = 1, then the velocity of the flow is equal to the velocity of surface waves, which is also known as the ...

  24. Numerical Analysis of Flow Characteristics and Energy ...

    The study utilizes laboratory experiments conducted in a flume measuring 12 m in length, 1.2 m in width, and 0.8 m in depth, with a chute angle of 26.6°. The cascade of spillways consists of 10 steps, each 0.12 m long and 0.06 m high. ... which proved helpful in energy dissipation. The downstream velocity and Froude number for case 4 were ...

  25. Existence and Stability of Nonmonotone Hydraulic Shocks for the Saint

    Extending the work of Yang-Zumbrun for the hydrodynamically stable case of Froude number \(F<2\), we categorize completely the existence and convective stability of hydraulic shock profiles of the Saint Venant equations of inclined thin film flow.Moreover, we confirm by numerical experiment that asymptotic dynamics for general Riemann data is given in the hydrodynamic instability regime by ...