Torricelli's law

Last updated
Torricelli's law describes the parting speed of a jet of water, based on the distance below the surface at which the jet starts, assuming no air resistance, viscosity, or other hindrance to the fluid flow. This diagram shows several such jets, vertically aligned, leaving the reservoir horizontally. In this case, the jets have an envelope (a concept also due to Torricelli) which is a line descending at 45deg from the water's surface over the jets. Each jet reaches farther than any other jet at the point where it touches the envelope, which is at twice the depth of the jet's source. The depth at which two jets cross is the sum of their source depths. Every jet (even if not leaving horizontally) takes a parabolic path whose directrix is the surface of the water. TorricelliLawDiagram.svg
Torricelli's law describes the parting speed of a jet of water, based on the distance below the surface at which the jet starts, assuming no air resistance, viscosity, or other hindrance to the fluid flow. This diagram shows several such jets, vertically aligned, leaving the reservoir horizontally. In this case, the jets have an envelope (a concept also due to Torricelli) which is a line descending at 45° from the water's surface over the jets. Each jet reaches farther than any other jet at the point where it touches the envelope, which is at twice the depth of the jet's source. The depth at which two jets cross is the sum of their source depths. Every jet (even if not leaving horizontally) takes a parabolic path whose directrix is the surface of the water.

Torricelli's law, also known as Torricelli's theorem, is a theorem in fluid dynamics relating the speed of fluid flowing from an orifice to the height of fluid above the opening. The law states that the speed of efflux of a fluid through a sharp-edged hole at the bottom of the tank filled to a depth is the same as the speed that a body would acquire in falling freely from a height , i.e. , where is the acceleration due to gravity. This expression comes from equating the kinetic energy gained, , with the potential energy lost, , and solving for . The law was discovered (though not in this form) by the Italian scientist Evangelista Torricelli, in 1643. It was later shown to be a particular case of Bernoulli's principle.

Contents

Derivation

Under the assumptions of an incompressible fluid with negligible viscosity, Bernoulli's principle states that the hydraulic energy is constant

at any two points in the flowing liquid. Here is fluid speed, is the acceleration due to gravity, is the height above some reference point, is the pressure, and is the density.

In order to derive Torricelli's formula the first point with no index is taken at the liquid's surface, and the second just outside the opening. Since the liquid is assumed to be incompressible, is equal to and; both can be represented by one symbol . The pressure and are typically both atmospheric pressure, so . Furthermore is equal to the height of the liquid's surface over the opening:

The velocity of the surface can by related to the outflow velocity by the continuity equation , where is the orifice's cross section and is the (cylindrical) vessel's cross section. Renaming to (A like Aperture) gives:

Torricelli's law is obtained as a special case when the opening is very small relative to the horizontal cross-section of the container :

Torricelli's law can only be applied when viscous effects can be neglected which is the case for water flowing out through orifices in vessels.

Experimental verification: Spouting can experiment

Experiment to determine the trajectory of an outflowing jet: Vertical rods are adjusted so they are nearly touching the jet. After the experiment the distance between a horizontal line and the location of the jet can be measured by the length adjustments of the rods. Experiment to determine the trajectory of an outflowing jet.jpg
Experiment to determine the trajectory of an outflowing jet: Vertical rods are adjusted so they are nearly touching the jet. After the experiment the distance between a horizontal line and the location of the jet can be measured by the length adjustments of the rods.

Every physical theory must be verified by experiments. The spouting can experiment consists of a cylindrical vessel filled up with water and with several holes in different heights. It is designed to show that in a liquid with an open surface, pressure increases with depth. The lower a jet is on the tube, the more powerful it is. The fluid exit velocity is greater further down the tube. [1]

The outflowing jet forms a downward parabola where every parabola reaches farther out the larger the distance between the orifice and the surface is. The shape of the parabola is only dependent on the outflow velocity and can be determined from the fact that every molecule of the liquid forms a ballistic trajectory (see projectile motion) where the initial velocity is the outflow velocity :

The results confirm the correctness of Torricelli's law very well.

Discharge and time to empty a cylindrical vessel

Assuming that a vessel is cylindrical with fixed cross-sectional area , with orifice of area at the bottom, then rate of change of water level height is not constant. The water volume in the vessel is changing due to the discharge out of the vessel:

Integrating both sides and re-arranging, we obtain

where is the initial height of the water level and is the total time taken to drain all the water and hence empty the vessel.

This formula has several implications. If a tank with volume with cross section and height , so that , is fully filled, then the time to drain all the water is

This implies that high tanks with same filling volume drain faster than wider ones.

Lastly, we can re-arrange the above equation to determine the height of the water level as a function of time as

where is the height of the container while is the discharge time as given above.

Discharge experiment, coefficient of discharge

The discharge theory can be tested by measuring the emptying time or time series of the water level within the cylindrical vessel. In many cases, such experiments do not confirm the presented discharge theory: when comparing the theoretical predictions of the discharge process with measurements, very large differences can be found in such cases. In reality, the tank usually drains much more slowly. Looking at the discharge formula

two quantities could be responsible for this discrepancy: the outflow velocity or the effective outflow cross section.

Figure 28 of Daniel Bernoulli's Hydrodynamica (1738) showing the generation of a vena contracta with streamlines. BernoulliHydrodynamicaFig28.jpg
Figure 28 of Daniel Bernoulli's Hydrodynamica (1738) showing the generation of a vena contracta with streamlines.

In 1738 Daniel Bernoulli attributed the discrepancy between the theoretical and the observed outflow behavior to the formation of a vena contracta which reduces the outflow cross-section from the orifice's cross-section to the contracted cross-section and stated that the discharge is:

Actually this is confirmed by state-of-the-art experiments (see [2] ) in which the discharge, the outflow velocity and the cross-section of the vena contracta were measured. Here it was also shown that the outflow velocity is predicted extremely well by Torricelli's law and that no velocity correction (like a "coefficient of velocity") is needed.

The problem remains how to determine the cross-section of the vena contracta. This is normally done by introducing a discharge coefficient which relates the discharge to the orifice's cross-section and Torricelli's law:

For low viscosity liquids (such as water) flowing out of a round hole in a tank, the discharge coefficient is in the order of 0.65. [3] By discharging through a round tube or hose, the coefficient of discharge can be increased to over 0.9. For rectangular openings, the discharge coefficient can be up to 0.67, depending on the height-width ratio.

Applications

Horizontal distance covered by the jet of liquid

If is height of the orifice above the ground and is height of the liquid column from the ground (height of liquid's surface), then the horizontal distance covered by the jet of liquid to reach the same level as the base of the liquid column can be easily derived. Since be the vertical height traveled by a particle of jet stream, we have from the laws of falling body

where is the time taken by the jet particle to fall from the orifice to the ground. If the horizontal efflux velocity is , then the horizontal distance traveled by the jet particle during the time duration is

Since the water level is above the orifice, the horizontal efflux velocity as given by Torricelli's law. Thus, we have from the two equations

The location of the orifice that yields the maximum horizontal range is obtained by differentiating the above equation for with respect to , and solving . Here we have

Solving we obtain

and the maximum range

Clepsydra problem

An inflow clepsydra Clepsydra.jpg
An inflow clepsydra

A clepsydra is a clock that measures time by the flow of water. It consists of a pot with a small hole at the bottom through which the water can escape. The amount of escaping water gives the measure of time. As given by the Torricelli's law, the rate of efflux through the hole depends on the height of the water; and as the water level diminishes, the discharge is not uniform. A simple solution is to keep the height of the water constant. This can be attained by letting a constant stream of water flow into the vessel, the overflow of which is allowed to escape from the top, from another hole. Thus having a constant height, the discharging water from the bottom can be collected in another cylindrical vessel with uniform graduation to measure time. This is an inflow clepsydra.

Alternatively, by carefully selecting the shape of the vessel, the water level in the vessel can be made to decrease at constant rate. By measuring the level of water remaining in the vessel, the time can be measured with uniform graduation. This is an example of outflow clepsydra. Since the water outflow rate is higher when the water level is higher (due to more pressure), the fluid's volume should be more than a simple cylinder when the water level is high. That is, the radius should be larger when the water level is higher. Let the radius increase with the height of the water level above the exit hole of area That is, . We want to find the radius such that the water level has a constant rate of decrease, i.e. .

At a given water level , the water surface area is . The instantaneous rate of change in water volume is

From Torricelli's law, the rate of outflow is

From these two equations,

Thus, the radius of the container should change in proportion to the quartic root of its height,

Likewise, if the shape of the vessel of the outflow clepsydra cannot be modified according to the above specification, then we need to use non-uniform graduation to measure time. The emptying time formula above tells us the time should be calibrated as the square root of the discharged water height, More precisely,

where is the time taken by the water level to fall from the height of to height of .

Torricelli's original derivation

Figures from Evangelista Torricelli's Opera Geometrica (1644) describing the derivation of his famous outflow formula: (a) One tube filled up with water from A to B. (b) In two connected tubes the water lift up to the same height. (c) When the tube C is removed, the water should rise up to the height D. Due to friction effects the water only rises to the point C. TorricelliOperaGeometricaOutflow.jpg
Figures from Evangelista Torricelli's Opera Geometrica (1644) describing the derivation of his famous outflow formula: (a) One tube filled up with water from A to B. (b) In two connected tubes the water lift up to the same height. (c) When the tube C is removed, the water should rise up to the height D. Due to friction effects the water only rises to the point C.

Evangelista Torricelli's original derivation can be found in the second book 'De motu aquarum' of his 'Opera Geometrica' (see [4] ): He starts a tube AB (Figure (a)) filled up with water to the level A. Then a narrow opening is drilled at the level of B and connected to a second vertical tube BC. Due to the hydrostatic principle of communicating vessels the water lifts up to the same filling level AC in both tubes (Figure (b)). When finally the tube BC is removed (Figure (c)) the water should again lift up to this height, which is named AD in Figure (c). The reason for that behavior is the fact that a droplet's falling velocity from a height A to B is equal to the initial velocity that is needed to lift up a droplet from B to A.

When performing such an experiment only the height C (instead of D in figure (c)) will be reached which contradicts the proposed theory. Torricelli attributes this defect to the air resistance and to the fact that the descending drops collide with ascending drops.

Torricelli's argumentation is, as a matter of fact, wrong because the pressure in free jet is the surrounding atmospheric pressure, while the pressure in a communicating vessel is the hydrostatic pressure. At that time the concept of pressure was unknown.

See also

Related Research Articles

<span class="mw-page-title-main">Pressure</span> Force distributed over an area

Pressure is the force applied perpendicular to the surface of an object per unit area over which that force is distributed. Gauge pressure is the pressure relative to the ambient pressure.

<span class="mw-page-title-main">Bernoulli's principle</span> Principle relating to fluid dynamics

Bernoulli's principle is a key concept in fluid dynamics that relates pressure, speed and height. Bernoulli's principle states that an increase in the speed of a fluid occurs simultaneously with a decrease in static pressure or the fluid's potential energy. The principle is named after the Swiss mathematician and physicist Daniel Bernoulli, who published it in his book Hydrodynamica in 1738. Although Bernoulli deduced that pressure decreases when the flow speed increases, it was Leonhard Euler in 1752 who derived Bernoulli's equation in its usual form.

<span class="mw-page-title-main">Speed of sound</span> Speed of sound wave through elastic medium

The speed of sound is the distance travelled per unit of time by a sound wave as it propagates through an elastic medium. At 20 °C (68 °F), the speed of sound in air is about 343 m/s, or one km in 2.91 s or one mile in 4.69 s. It depends strongly on temperature as well as the medium through which a sound wave is propagating. At 0 °C (32 °F), the speed of sound in air is about 331 m/s. More simply, the speed of sound is how fast vibrations travel.

In fluid dynamics, the Darcy–Weisbach equation is an empirical equation that relates the head loss, or pressure loss, due to friction along a given length of pipe to the average velocity of the fluid flow for an incompressible fluid. The equation is named after Henry Darcy and Julius Weisbach. Currently, there is no formula more accurate or universally applicable than the Darcy-Weisbach supplemented by the Moody diagram or Colebrook equation.

<span class="mw-page-title-main">Terminal velocity</span> Highest velocity attainable by a falling object

Terminal velocity is the maximum velocity (speed) attainable by an object as it falls through a fluid. It occurs when the sum of the drag force (Fd) and the buoyancy is equal to the downward force of gravity (FG) acting on the object. Since the net force on the object is zero, the object has zero acceleration. For objects falling through regular air, the buoyant force is usually dismissed and not taken into account as its effects are negligible.

A Newtonian fluid is a fluid in which the viscous stresses arising from its flow are at every point linearly correlated to the local strain rate — the rate of change of its deformation over time. Stresses are proportional to the rate of change of the fluid's velocity vector.

In continuum mechanics, the Froude number is a dimensionless number defined as the ratio of the flow inertia to the external field. The Froude number is based on the speed–length ratio which he defined as:

<span class="mw-page-title-main">Siphon</span> Device involving the flow of liquids through tubes

A siphon is any of a wide variety of devices that involve the flow of liquids through tubes. In a narrower sense, the word refers particularly to a tube in an inverted "U" shape, which causes a liquid to flow upward, above the surface of a reservoir, with no pump, but powered by the fall of the liquid as it flows down the tube under the pull of gravity, then discharging at a level lower than the surface of the reservoir from which it came.

<span class="mw-page-title-main">Hydrostatics</span> Branch of fluid mechanics that studies fluids at rest

Fluid statics or hydrostatics is the branch of fluid mechanics that studies fluids at hydrostatic equilibrium and "the pressure in a fluid or exerted by a fluid on an immersed body".

<span class="mw-page-title-main">Venturi effect</span> Reduced pressure caused by a flow restriction in a tube or pipe

The Venturi effect is the reduction in fluid pressure that results when a moving fluid speeds up as it flows through a constricted section of a pipe. The Venturi effect is named after its discoverer, the 18th-century Italian physicist Giovanni Battista Venturi.

An orifice plate is a device used for measuring flow rate, for reducing pressure or for restricting flow.

<span class="mw-page-title-main">Hydraulic head</span> Specific measurement of liquid pressure above a vertical datum

Hydraulic head or piezometric head is a specific measurement of liquid pressure above a vertical datum.

Fluid mechanics is the branch of physics concerned with the mechanics of fluids and the forces on them. It has applications in a wide range of disciplines, including mechanical, aerospace, civil, chemical, and biomedical engineering, as well as geophysics, oceanography, meteorology, astrophysics, and biology.

Choked flow is a compressible flow effect. The parameter that becomes "choked" or "limited" is the fluid velocity.

Accidental release source terms are the mathematical equations that quantify the flow rate at which accidental releases of liquid or gaseous pollutants into the ambient environment which can occur at industrial facilities such as petroleum refineries, petrochemical plants, natural gas processing plants, oil and gas transportation pipelines, chemical plants, and many other industrial activities. Governmental regulations in many countries require that the probability of such accidental releases be analyzed and their quantitative impact upon the environment and human health be determined so that mitigating steps can be planned and implemented.

The flow coefficient of a device is a relative measure of its efficiency at allowing fluid flow. It describes the relationship between the pressure drop across an orifice valve or other assembly and the corresponding flow rate.

<span class="mw-page-title-main">Shallow water equations</span> Set of partial differential equations that describe the flow below a pressure surface in a fluid

The shallow-water equations (SWE) are a set of hyperbolic partial differential equations that describe the flow below a pressure surface in a fluid. The shallow-water equations in unidirectional form are also called Saint-Venant equations, after Adhémar Jean Claude Barré de Saint-Venant.

In nonideal fluid dynamics, the Hagen–Poiseuille equation, also known as the Hagen–Poiseuille law, Poiseuille law or Poiseuille equation, is a physical law that gives the pressure drop in an incompressible and Newtonian fluid in laminar flow flowing through a long cylindrical pipe of constant cross section. It can be successfully applied to air flow in lung alveoli, or the flow through a drinking straw or through a hypodermic needle. It was experimentally derived independently by Jean Léonard Marie Poiseuille in 1838 and Gotthilf Heinrich Ludwig Hagen, and published by Hagen in 1839 and then by Poiseuille in 1840–41 and 1846. The theoretical justification of the Poiseuille law was given by George Stokes in 1845.

In a nozzle or other constriction, the discharge coefficient is the ratio of the actual discharge to the ideal discharge, i.e., the ratio of the mass flow rate at the discharge end of the nozzle to that of an ideal nozzle which expands an identical working fluid from the same initial conditions to the same exit pressures.

<span class="mw-page-title-main">Stokes problem</span>

In fluid dynamics, Stokes problem also known as Stokes second problem or sometimes referred to as Stokes boundary layer or Oscillating boundary layer is a problem of determining the flow created by an oscillating solid surface, named after Sir George Stokes. This is considered one of the simplest unsteady problems that has an exact solution for the Navier–Stokes equations. In turbulent flow, this is still named a Stokes boundary layer, but now one has to rely on experiments, numerical simulations or approximate methods in order to obtain useful information on the flow.

References

  1. Spouting cylinder fluid flow.
  2. J.H. Lienhard (V) and J.H. Lienhard (IV): Velocity Coefficients for Free Jets from Sharp-Edged Orifices, Journal of Fluids Engineering 106,13-17,1984, https://doi.org/10.1115/1.3242391
  3. tec-science (2019-11-21). "Discharge of liquids (Torricelli's law)". tec-science. Retrieved 2019-12-08.
  4. A. Malcherek: History of the Torricelli Principle and a New Outflow Theory, Journal of Hydraulic Engineering 142(11),1-7,2016, https://doi.org/10.1061/(ASCE)HY.1943-7900.0001232)

Further reading