Anda di halaman 1dari 51

Darcy's law:

In fluid dynamics and hydrology, Darcy's law is a phenomenologically derived constitutive


equation that describes the flow of a fluid through a porous medium. The law was formulated
by Henry Darcy based on the results of experiments on the flow of water through beds of
sand. It also forms the scientific basis of fluid permeability used in the earth sciences.

Background:

Although Darcy's law (an expression of conservation of momentum) was determined


experimentally by Henri Darcy, it has since been derived from the Navier-Stokes equations
via homogenization. It is analogous to Fourier's law in the field of heat conduction, Ohm's
law in the field of electrical networks, or Fick's law in diffusion theory.

One application of Darcy's law is water flow through an aquifer. Darcy's law along with the
equation of conservation of mass both are equivalent to the groundwater flow equation, one
of the basic relationships of hydrogeology. Darcy's law is also used to describe oil, water, and
gas flows through petroleum reservoirs.

Description:

Diagram showing definitions and directions for Darcy's law.

Darcy's law is a simple proportional relationship between the instantaneous discharge rate
through a porous medium, the viscosity of the fluid and the pressure drop over a given
distance.

The total discharge, Q (units of volume per time, e.g., m³/s) is equal to the product of the
permeability (κ units of area, e.g. m²) of the medium, the cross-sectional area (A) to flow, and
the pressure drop (Pb − Pa), all divided by the dynamic viscosity μ (in SI units e.g. kg/(m·s) or
Pa·s), and the length L the pressure drop is taking place over. The negative sign is needed
because fluids flow from high pressure to low pressure. So if the change in pressure is
negative (in the x-direction) then the flow will be positive (in the x-direction). Dividing both
sides of the equation by the area and using more general notation leads to

1
Where q is the flux (discharge per unit area, with units of length per time, m/s) and is the
pressure gradient vector. This value of flux, often referred to as the Darcy flux, is not the
velocity which the water traveling through the pores is experiencing.

The pore velocity (v) is related to the Darcy flux (q) by the porosity (φ). The flux is divided
by porosity to account for the fact that only a fraction of the total formation volume is
available for flow. The pore velocity would be the velocity a conservative tracer would
experience if carried by the fluid through the formation.

In 3D:

In three dimensions, gravity must be accounted for, as the flow is not affected by the vertical
pressure drop caused by gravity when assuming hydrostatic conditions. The solution is to
subtract the gravitational pressure drop from the existing pressure drop in order to express the
resulting flow,

where the flux is now a vector quantity, is a tensor of permeability, is the gradient
operator in 3D, g is the acceleration due to gravity, is the unit vector in the vertical
direction, pointing downwards and ρ is the density.

Effects of anisotropy in three dimensions are addressed using a symmetric second-order


tensor of permeability:

where the magnitudes of permeability in the x, y, z component directions are specified. Since
this a symmetric matrix, there are at most six unique values. If the permeability is isotropic
(equal magnitude in all directions), then the diagonal values are equal,
, while all other components are 0. The permeability tensor can be
interpreted through an evaluation of the relative magnitudes of each component. For example,
rock with highly permeable vertical fractures aligned in the x-direction will have relatively
higher values for than other component values.

Assumptions:

Darcy's law is a simple mathematical statement which neatly summarizes several familiar
properties that groundwater flowing in aquifers exhibits, including:
2
• If there is no pressure gradient over a distance, no flow occurs (this is hydrostatic
conditions).
• If there is a pressure gradient, flow will occur from high pressure towards low
pressure (opposite the direction of increasing gradient - hence the negative sign in
Darcy's law).
• The greater the pressure gradient (through the same formation material), the greater
the discharge rate, and
• The discharge rate of fluid will often be different — through different formation
materials (or even through the same material, in a different direction) — even if the
same pressure gradient exists in both cases.

A graphical illustration of the use of the steady-state groundwater flow equation (based on
Darcy's law and the conservation of mass) is in the construction of flownets, to quantify the
amount of groundwater flowing under a dam.

Darcy's law is only valid for slow, viscous flow; fortunately, most groundwater flow cases
fall in this category. Typically any flow with a Reynolds number less than one is clearly
laminar, and it would be valid to apply Darcy's law. Experimental tests have shown that flow
regimes with values of Reynolds number up to 10 may still be Darcian. Reynolds number (a
dimensionless parameter) for porous media flow is typically expressed as

where ρ is the density of the fluid (units of mass per volume), v is the specific discharge (not
the pore velocity — with units of length per time), d30 is a representative grain diameter for
the porous medium (often taken as the 30% passing size from a grain size analysis using
sieves), and μ is the dynamic viscosity of the fluid.

Derivation:

Assuming stationary, creeping, incompressible flow, the Navier-Stokes equation becomes

Where μ is the viscosity, ui is the velocity in the i direction, gi is the gravity component in the
i direction and p is the pressure. Assuming the viscous resisting force is proportional to the
velocity, and opposite in direction, we may write

where φ is the porosity. This gives the velocity

3
which gives Darcy's law

Additional forms of Darcy's law:

Time derivative of flux:

For very short time scales, a time derivative of flux may be added to Darcy's law, which
results in valid solutions at very small times (in heat transfer, this is called the modified form
of Fourier's law),

where τ is a very small time constant which causes this equation to reduce to the normal form
of Darcy's law at "normal" times (> nanoseconds). The main reason for doing this is that the
regular groundwater flow equation (diffusion equation) leads to singularities at constant head
boundaries at very small times. This form is more mathematically rigorous, but leads to a
hyperbolic groundwater flow equation, which is more difficult to solve and is only useful at
very small times, typically out of the realm of practical use.

Brinkman term:

Another extension to the traditional form of Darcy's law is the Brinkman term, which is used
to account for transitional flow between boundaries (introduced by Brinkman in 1947),

where β is an effective viscosity term. This correction term accounts for flow through
medium where the grains of the media are porous themselves, but is difficult to use, and is
typically neglected.

Multiphase flow:

For multiphase flow, an approximation is to use Darcy's law for each phase, with
permeability replaced by phase permeability, which is the permeability of the rock multiplied
with relative permeability. This approximation is valid if the interfaces between the fluids
remain static, which is not true in general, but it is still a reasonable model under steady-state
conditions.

Assuming that the flow of a phase in the presence of another phase can be viewed as single
phase flow through a reduced pore network, we can add the subscript i for each phase to
Darcy's law above written for Darcy flux, and obtain for each phase in multiphase flow

4
where κi is the phase permeability for phase i. From this we also define relative permeability
κri for phase i as

κri = κi / κ

Where κ is the permeability for the porous medium as in Darcy's law

Forchheimer equation for non-Darcy flow:

For a sufficiently high flow velocity, the flow is nonlinear, and Dupuit and Forchheimer has
proposed to generalize the flow equation to

Where V is the flow velocity and β is a factor to be experimentally deduced.

In membrane operations:

In pressure-driven membrane operations, Darcy's law is often used in the form,

Where,

• J is the volumetric flux (m.s − 1),


• ΔP is the hydraulic pressure difference between the feed and permeate sides of the
membrane (Pa),
• ΔΠ is the osmotic pressure difference between the feed and permeate sides of the
membrane (Pa),
• μ is the dynamic viscosity (Pa.s),
• Rf is the fouling resistance (m − 1), and
• Rm is the membrane resistance (m − 1).

Darcy–Weisbach equation:
In fluid dynamics, the Darcy–Weisbach equation is a phenomenological equation, which
relates the head loss — or pressure loss — due to friction along a given length of pipe to the
average velocity of the fluid flow. The equation is named after Henry Darcy and Julius
Weisbach.

The Darcy–Weisbach equation contains a dimensionless friction factor, known as the Darcy
friction factor. This is also called the Darcy–Weisbach friction factor or Moody friction

5
factor. The Darcy friction factor is four times the Fanning friction factor, with which it
should not be confused.

Head loss form:

Head loss can be calculated with

Where

• hf is the head loss due to friction;


• L is the length of the pipe;
• D is the hydraulic diameter of the pipe (for a pipe of circular section, this equals the
internal diameter of the pipe);
• V is the average velocity of the fluid flow, equal to the volumetric flow rate per unit
cross-sectional wetted area;
• g is the local acceleration due to gravity;
• f is a dimensionless coefficient called the Darcy friction factor. It can be found from a
Moody diagram.

Pressure loss form:

Given that the head loss hf expresses the pressure loss Δp as the height of a column of fluid,

Where ρ is the density of the fluid, the Darcy–Weisbach equation can also be written in terms
of pressure loss:

Where the pressure loss due to friction Δp is a function of:

• the ratio of the length to diameter of the pipe, L/D;


• the density of the fluid, ρ;
• the average velocity of the flow, V, as defined above;
• a (dimensionless) coefficient of laminar, or turbulent flow, f.

Since the pressure loss equation can be derived from the head loss equation by multiplying
each side by ρ and g.

6
Darcy friction factor:

The friction factor f or flow coefficient λ is not a constant and depends on the parameters of
the pipe and the velocity of the fluid flow, but it is known to high accuracy within certain
flow regimes. It may be evaluated for given conditions by the use of various empirical or
theoretical relations, or it may be obtained from published charts. These charts are often
referred to as Moody diagrams, after L. F. Moody, and hence the factor itself is sometimes
called the Moody friction factor. It is also sometimes called the Blasius friction factor, after
the approximate formula he proposed.

For laminar (slow) flows, it is a consequence of Poiseuille's Law that λ equals 64/Re, where
Re is the Reynolds Number calculated substituting for the characteristic length the hydraulic
diameter of the pipe, which equals the inside diameter for circular pipe geometries.

For turbulent flow, methods for finding the friction factor f include using a diagram such as
the Moody chart; or solving equations such as the Colebrook-White equation, or the Swamee-
Jain equation. While the diagram and Colebrook-White equation are iterative methods, the
Swamee-Jain equation allows f to be found directly for full flow in a circular pipe.

Confusion with the Fanning friction factor:

The Darcy–Weisbach friction factor is 4 times larger than the Fanning friction factor, so
attention must be paid to note which one of these is meant in any "friction factor" chart or
equation being used. Of the two, the Darcy–Weisbach factor is more commonly used by civil
and mechanical engineers and the Fanning factor by chemical engineers, but care should be
taken to identify the correct factor regardless of the source of the chart or formula.

Most charts or tables indicate the type of friction factor, or at least provide the formula for the
friction factor with laminar flow.

If the formula is f=16/Re, it's the Fanning factor, and if the formula is f=64/Re, it's the
Darcy–Weisbach factor.

Which friction factor is plotted in a Moody diagram may be determined by inspection if the
publisher did not include the formula described above:

1. Observe the value of the friction factor for laminar flow at a Reynolds number of
1000.
2. If the value of the friction factor is 0.064, then the Darcy friction factor is plotted in
the Moody diagram. Note that the nonzero digits in 0.064 are the numerator in the
formula for the laminar Darcy friction factor: f = 64/Re.
3. If the value of the friction factor is 0.016, then the Fanning friction factor is plotted in
the Moody diagram. Note that the nonzero digits in 0.016 are the numerator in the
formula for the laminar Fanning friction factor: f = 16/Re.

The procedure above is similar for any available Reynolds number that is an integral power
of ten. It is not necessary to remember the value 1000 for this procedure – only that an
integral power of ten is of interest for this purpose.

7
History:

Historically this equation arose as a variant on the Prony equation; this variant was developed
by Henry Darcy of France, and further refined into the form used today by Julius Weisbach
of Saxony in 1845. Initially, data on the variation of f with velocity was lacking, so the
Darcy–Weisbach equation was outperformed at first by the empirical Prony equation in many
cases. In later years it was eschewed in many special-case situations in favor of a variety of
empirical equations valid only for certain flow regimes, notably the Hazen-Williams equation
or the Manning equation, most of which were significantly easier to use in calculations.
However, since the advent of the calculator, ease of calculation is no longer a major issue,
and so the Darcy–Weisbach equation's generality has made it the preferred one.

Derivation:

The Darcy–Weisbach equation is a phenomenological formula obtainable by dimensional


analysis.

Away from the ends of the pipe, the characteristics of the flow are independent of the
position along the pipe. The key quantities are then the pressure drop along the pipe per unit
length, Δp/L, and the volumetric flow rate. The flow rate can be converted to an average
velocity V by dividing by the wetted area of the flow (which equals the cross-sectional area of
the pipe if the pipe is full of fluid).

Pressure has dimensions of energy per unit volume and, so, the pressure drop between two
points must be proportional to (1/2)ρV2 which has the same dimensions as it resembles (see
below) the expression for the kinetic energy per unit volume. We also know that it must be
proportional to the length of the pipe between the two points L as the pressure drop per unit
length is a constant. To turn that into a dimensionless quantity we can divide by the hydraulic
diameter of the pipe D which is also constant along the pipe. Therefore,

The proportionality coefficient is the dimensionless "Darcy friction factor" or "flow


coefficient". This dimensionless coefficient will be a combination of geometric factors such
as π, the Reynolds number and (outside the laminar regime) the relative roughness of the pipe
(the ratio of the roughness height to the hydraulic diameter).

Note that (1/2)ρV2 is not the kinetic energy of the fluid per unit volume, for the following
reasons. Even in the case of laminar flow, where all the flow lines are parallel to the length of
the pipe, the velocity of the fluid on the inner surface of the pipe is zero due to viscosity, and
the velocity in the center of the pipe must therefore be larger than the average velocity
obtained by dividing the volumetric flow rate by the wet area. The average kinetic energy
then involves the mean-square velocity, which always exceeds the square of the mean
velocity. In the case of turbulent flow, the fluid acquires random velocity components in all
directions, including perpendicular to the length of the pipe, and thus turbulence contributes
to the kinetic energy per unit volume but not to the average lengthwise velocity of the fluid.

8
Bernoulli's principle:
This article is about Bernoulli's principle and Bernoulli's equation in fluid dynamics. For an
unrelated topic in ordinary differential equations, see Bernoulli differential equation .

The flow of air into a venturi meter. The kinetic energy increases at the expense of the fluid
pressure, as shown by the difference in height of the two columns of water.

In fluid dynamics, Bernoulli's principle states that for an inviscid flow, an increase in the
speed of the fluid occurs simultaneously with a decrease in pressure or a decrease in the
fluid's potential energy. Bernoulli's principle is named after the Dutch–Swiss mathematician
Daniel Bernoulli who published his principle in his book Hydrodynamica in 1738.

Bernoulli's principle can be applied to various types of fluid flow, resulting in what is loosely
denoted as Bernoulli's equation. In fact, there are different forms of the Bernoulli equation
for different types of flow. The simple form of Bernoulli's principle is valid for
incompressible flows (e.g. most liquid flows) and also for compressible flows (e.g. gases)
moving at low Mach numbers. More advanced forms may in some cases be applied to
compressible flows at higher Mach numbers (see the derivations of the Bernoulli equation).

Bernoulli's principle is equivalent to the principle of conservation of energy. This states


that in a steady flow the sum of all forms of mechanical energy in a fluid along a streamline
is the same at all points on that streamline. This requires that the sum of kinetic energy and
potential energy remain constant. If the fluid is flowing out of a reservoir the sum of all forms
of energy is the same on all streamlines because in a reservoir the energy per unit mass (the
sum of pressure and gravitational potential ρgh) is the same everywhere.

Fluid particles are subject only to pressure and their own weight. If a fluid is flowing
horizontally and along a section of a streamline, where the speed increases it can only be
because the fluid on that section has moved from a region of higher pressure to a region of
lower pressure; and if its speed decreases, it can only be because it has moved from a region
of lower pressure to a region of higher pressure. Consequently, within a fluid flowing
horizontally, the highest speed occurs where the pressure is lowest, and the lowest speed
occurs where the pressure is highest

Incompressible flow equation:

In most flows of liquids, and of gases at low Mach number, the mass density of a fluid parcel
can be considered to be constant, regardless of pressure variations in the flow. For this reason
the fluid in such flows can be considered to be incompressible and these flows can be

9
described as incompressible flow. Bernoulli performed his experiments on liquids and his
equation in its original form is valid only for incompressible flow.

The original form of Bernoulli's equation — valid at any point along a streamline — is:

Where:

Is the fluid flow speed at a point on a streamline.

Is the acceleration due to gravity.

Is the elevation of the point above a reference plane, with the positive z-direction
pointing upward — so in the direction opposite to the gravitational acceleration.

Is the pressure at the point, and

Is the density of the fluid at all points in the fluid.

The following assumptions must be met for the equation to apply:

• The fluid must be incompressible - even though pressure varies, the density must
remain constant.
• The streamline must not enter a boundary layer. (Bernoulli's equation is not applicable
where there are viscous forces, such as in a boundary layer.)

By multiplying with the mass density ( ), the above equation can be rewritten as:

Or:

Where:

is dynamic pressure,

is the piezometric head or hydraulic head (the sum of the elevation z


and the pressure head) and

Is the total pressure (the sum of the static pressure p and dynamic
pressure q).

10
The constant in the Bernoulli equation can be normalized. A common approach is in terms of
total head or energy head H:

so divide the above constant by ρ and g to get the total head H in terms of meters of
fluid column.

The above equations suggest there is a flow speed at which pressure is zero, and at even
higher speeds the pressure is negative. Most often, gases and liquids are not capable of
negative absolute pressure, or even zero pressure, so clearly Bernoulli's equation ceases to be
valid before zero pressure is reached. In liquids when the pressure becomes too low
cavitation occurs. The above equations use a linear relationship between flow speed squared
and pressure. At higher flow speeds in gases, or for sound waves in liquids, the changes in
mass density become significant so that the assumption of constant density is invalid.

Simplified form:

In many applications of Bernoulli's equation, the change in the term along the streamline
is so small compared with the other terms it can be ignored. For example, in the case of
aircraft in flight, the change in height along a streamline is so small the term can be
omitted. This allows the above equation to be presented in the following simplified form:

Where is called total pressure and dynamic pressure. Many authors refer to the pressure
as static pressure to distinguish it from total pressure and dynamic pressure . In
Aerodynamics, L.J. Clancy writes: "To distinguish it from the total and dynamic pressures,
the actual pressure of the fluid, which is associated not with its motion but with its state, is
often referred to as the static pressure, but where the term pressure alone is used it refers to
this static pressure."

The simplified form of Bernoulli's equation can be summarized in the following memorable
word equation:

static pressure + dynamic pressure = total pressure[10]

Every point in a steadily flowing fluid, regardless of the fluid speed at that point, has its own
unique static pressure and dynamic pressure . Their sum is defined to be the total
pressure . The significance of Bernoulli's principle can now be summarized as total
pressure is constant along a streamline.

If the fluid flow is irrotational, the total pressure on every streamline is the same and
Bernoulli's principle can be summarized as total pressure is constant everywhere in the fluid
flow. It is reasonable to assume that irrotational flow exists in any situation where a large
body of fluid is flowing past a solid body. Examples are aircraft in flight, and ships moving in
open bodies of water. However, it is important to remember that Bernoulli's principle does
not apply in the boundary layer or in fluid flow through long pipes.
11
Applicability of incompressible flow equation to flow of
gases:

Bernoulli's equation is sometimes valid for the flow of gases: provided that there is no
transfer of kinetic or potential energy from the gas flow to the compression or expansion of
the gas. If both the gas pressure and volume change simultaneously, then work will be done
on or by the gas. In this case, Bernoulli's equation — in its incompressible flow form —
cannot be assumed to be valid. However if the gas process is entirely isobaric or isochoric
then no work is done on or by the gas (so the simple energy balance is not upset). According
to the gas law, an isobaric or isochoric process is ordinarily the only way to ensure constant
density in a gas. Also the gas density will be proportional to the ratio of pressure and absolute
temperature, however this ratio will vary upon compression or expansion, no matter what
non-zero quantity of heat is added or removed. The only exception is if the net heat transfer is
zero, as in a complete thermodynamic cycle, or in an individual isentropic (frictionless
adiabatic) process, and even then this reversible process must be reversed, to restore the gas
to the original pressure and specific volume, and thus density. Only then is the original,
unmodified Bernoulli equation applicable. In this case the equation can be used if the flow
speed of the gas is sufficiently below the speed of sound, such that the variation in density of
the gas (due to this effect) along each streamline can be ignored. Adiabatic flow at less than
Mach 0.3 is generally considered to be slow enough.

Unsteady potential flow:

The Bernoulli equation for unsteady potential flow is used in the theory of ocean surface
waves and acoustics.

For an irrotational flow, the flow velocity can be described as the gradient ∇φ of a velocity
potential φ. In that case, and for a constant density ρ, the momentum equations of the Euler
equations can be integrated to:[12]

Which is a Bernoulli equation valid also for unsteady or time dependent flow. Here ∂φ/∂t
denotes the partial derivative of the velocity potential φ with respect to time t, and v = |∇φ| is
the flow speed. The function f(t) depends only on time and not on position in the fluid. As a
result, the Bernoulli equation at some moment t does not only apply along a certain
streamline, but in the whole fluid domain. This is also true for the special case of a steady
irrotational flow, in which case f is a constant.

Further f(t) can be made equal to zero by incorporating it into the velocity potential using the
transformation

Resulting in

Note that the relation of the potential to the flow velocity is unaffected by this transformation:
∇Φ = ∇φ.
12
The Bernoulli equation for unsteady potential flow also appears to play a central role in
Luke's variational principle, a variational description of free-surface flows using the
Lagrangian (not to be confused with Lagrangian coordinates).

Compressible flow equation:

Bernoulli developed his principle from his observations on liquids, and his equation is
applicable only to incompressible fluids, and compressible fluids at very low speeds (perhaps
up to 1/3 of the sound speed in the fluid). It is possible to use the fundamental principles of
physics to develop similar equations applicable to compressible fluids. There are numerous
equations, each tailored for a particular application, but all are analogous to Bernoulli's
equation and all rely on nothing more than the fundamental principles of physics such as
Newton's laws of motion or the first law of thermodynamics.

Compressible flow in fluid dynamics:

A useful form of the equation, suitable for use in compressible fluid dynamics, is:

(Constant along a streamline)

Where:

Is the ratio of the specific heats of the fluid

Is the pressure at a point

Is the density at the point

Is the speed of the fluid at the point

Is the acceleration due to gravity

Is the elevation of the point above a reference plane

In many applications of compressible flow, changes in elevation are negligible compared to


the other terms, so the term can be omitted. A very useful form of the equation is then:

Where:

Is the total pressure

Is the total density

Compressible flow in thermodynamics:


13
Another useful form of the equation, suitable for use in thermodynamics, is:

is the enthalpy per unit mass, which is also often written as (not to be confused with
"head" or "height").

Note that

Where is the thermodynamic energy per unit mass, also known as the specific internal
energy or "sie."

The constant on the right hand side is often called the Bernoulli constant and denoted . For
steady inviscid adiabatic flow with no additional sources or sinks of energy, is constant
along any given streamline. More generally, when may vary along streamlines, it still
proves a useful parameter, related to the "head" of the fluid (see below).

When the change in can be ignored, a very useful form of this equation is:

Where is total enthalpy

When shock waves are present, in a reference frame moving with a shock, many of the
parameters in the Bernoulli equation suffer abrupt changes in passing through the shock. The
Bernoulli parameter itself, however, remains unaffected. An exception to this rule is radiative
shocks, which violate the assumptions leading to the Bernoulli equation, namely the lack of
additional sinks or sources of energy.

Derivations of Bernoulli equation:

Real world application:

In modern every-day life there are many observations that can be successfully explained by
application of Bernoulli's principle.

• The relative air flow parallel to the top surface of an aircraft wing or helicopter rotor
blade is faster than along the bottom surface. Bernoulli's principle states that the
pressure on the surfaces of the wing or rotor blade will be lower above than below,
and this pressure difference results in an upwards lift force. If the relative air flows
across the top and bottom surfaces of a wing or rotor are known, then lift forces can
be calculated (to a good approximation) using Bernoulli's equations — established by
Bernoulli over a century before the first man-made wings were used for the purpose
of flight. Note that Bernoulli's principle does not explain why the air flows faster past
the top of the wing and slower past the under-side. To understand why, it is helpful to
14
understand circulation, the Kutta condition and the Kutta–Joukowski theorem.
Besides, Newton's third law states that forces only exist in pairs, so the air's upwards
force on the wing coexists with the wing's downward force on the air, which results in
a downward acceleration of air.

• The carburetor used in many reciprocating engines contains a venturi to create a


region of low pressure to draw fuel into the carburetor and mix it thoroughly with the
incoming air. The low pressure in the throat of a venturi can be explained by
Bernoulli's principle - in the narrow throat, the air is moving at its fastest speed and
therefore it is at its lowest pressure.

• The Pitot tube and static port on an aircraft are used to determine the airspeed of the
aircraft. These two devices are connected to the airspeed indicator which determines
the dynamic pressure of the airflow past the aircraft. Dynamic pressure is the
difference between stagnation pressure and static pressure. Bernoulli's principle is
used to calibrate the airspeed indicator so that it displays the indicated airspeed
appropriate to the dynamic pressure.

• The flow speed of a fluid can be measured using a device such as a Venturi meter or
an orifice plate, which can be placed into a pipeline to reduce the diameter of the
flow. For a horizontal device, the continuity equation shows that for an
incompressible fluid, the reduction in diameter will cause an increase in the fluid flow
speed. Subsequently Bernoulli's principle then shows that there must be a decrease in
the pressure in the reduced diameter region. This phenomenon is known as the
Venturi effect.

• The maximum possible drain rate for a tank with a hole or tap at the base can be
calculated directly from Bernoulli's equation, and is found to be proportional to the
square root of the height of the fluid in the tank. This is Torricelli's law, showing that
Torricelli's law is compatible with Bernoulli's principle. Viscosity lowers this drain
rate. This is reflected in the discharge coefficient which is a function of the Reynold's
number and the shape of the orifice.

• The principle also makes it possible for sail-powered craft to travel faster than the
wind that propels them (if friction can be sufficiently reduced). If the wind passing in
front of the sail is fast enough to experience a significant reduction in pressure, the
sail is pulled forward, in addition to being pushed from behind. Although boats in
water must contend with the friction of the water along the hull, ice sailing and land
sailing vehicles can travel faster than the wind.

Misunderstandings about the generation of lift:

Many explanations for the generation of lift (on airfoils, propeller blades, etc) can be found;
but some of these explanations can be misleading, and some are false. This has been a source
of heated discussion over the years. In particular, there has been debate about whether lift is
best explained by Bernoulli's principle or Newton's laws of motion. Modern writings agree
that Bernoulli's principle and Newton's laws are both relevant and correct.

Several of these explanations use Bernoulli's principle to connect the flow kinematics to the
flow-induced pressures. In case of incorrect (or partially correct) explanations of lift, also
15
relying at some stage on Bernoulli's principle, the errors generally occur in the assumptions
on the flow kinematics, and how these are produced. It is not Bernoulli's principle itself that
is questioned because this principle is well established.

Navier–Stokes equations
Continuum mechanics

This box: view • talk • edit

The Navier–Stokes equations, named after Claude-Louis Navier and George Gabriel Stokes,
describe the motion of fluid substances, that is substances which can flow. These equations
arise from applying Newton's second law to fluid motion, together with the assumption that
the fluid stress is the sum of a diffusing viscous term (proportional to the gradient of
velocity), plus a pressure term.

They are one of the most useful sets of equations because they describe the physics of a large
number of phenomena of academic and economic interest. They may be used to model
weather, ocean currents, water flow in a pipe, flow around an airfoil (wing), and motion of
stars inside a galaxy. As such, these equations in both full and simplified forms are used in
the design of aircraft and cars, the study of blood flow, the design of power stations, the
analysis of the effects of pollution, etc. Coupled with Maxwell's equations they can be used to
model and study magnetohydrodynamics.

The Navier–Stokes equations are also of great interest in a purely mathematical sense.
Somewhat surprisingly, given their wide range of practical uses, mathematicians have not yet
proven that in three dimensions solutions always exist (existence), or that if they do exist they
do not contain any infinities, singularities or discontinuities (smoothness). These are called
the Navier–Stokes existence and smoothness problems. The Clay Mathematics Institute has
called this one of the seven most important open problems in mathematics, and offered a
US$1,000,000 prize for a solution or a counter-example.

The Navier–Stokes equations are differential equations which, unlike algebraic equations, do
not explicitly establish a relation among the variables of interest (e.g. velocity and pressure).
Rather, they establish relations among the rates of change. For example, the Navier–Stokes
equations for simple case of an ideal fluid (inviscid and incompressible) can state that

16
acceleration (the rate of change of velocity) is proportional to the gradient (a type of
multivariate derivative) of pressure.

The Navier–Stokes equations dictate not position but rather velocity. A solution of the
Navier–Stokes equations is called a velocity field or flow field, which is a description of the
velocity of the fluid at a given point in space and time. Once the velocity field is solved for,
other quantities of interest (such as flow rate or drag force) may be found. This is different
from what one normally sees in classical mechanics, where solutions are typically trajectories
of position of a particle or deflection of a continuum. Studying velocity instead of position
makes more sense for a fluid; however for visualization purposes one can compute various
trajectories.

Properties:

Nonlinearity:

The Navier–Stokes equations are nonlinear partial differential equations in almost every real
situation. In some cases, such as one-dimensional flow and Stokes flow (or creeping flow),
the equations can be simplified to linear equations. The nonlinearity makes most problems
difficult or impossible to solve and is the main contributor to the turbulence that the equations
model.

The nonlinearity is due to convective acceleration, which is an acceleration associated with


the change in velocity over position. Hence, any convective flow, whether turbulent or not,
will involve nonlinearity, an example of convective but laminar (nonturbulent) flow would be
the passage of a viscous fluid (for example, oil) through a small converging nozzle. Such
flows, whether exactly solvable or not, can often be thoroughly studied and understood.

Turbulence:

Turbulence is the time dependent chaotic behavior seen in many fluid flows. It is generally
believed that it is due to the inertia of the fluid as a whole: the culmination of time dependent
and convective acceleration; hence flows where inertial effects are small tend to be laminar
(the Reynolds number quantifies how much the flow is affected by inertia). It is believed,
though not known with certainty, that the Navier–Stokes equations describe turbulence
properly.

The numerical solution of the Navier–Stokes equations for turbulent flow is extremely
difficult, and due to the significantly different mixing-length scales that are involved in
turbulent flow, the stable solution of this requires such a fine mesh resolution that the
computational time becomes significantly infeasible for calculation (see Direct numerical
simulation). Attempts to solve turbulent flow using a laminar solver typically result in a time-
unsteady solution, which fails to converge appropriately. To counter this, time-averaged
equations such as the Reynolds-averaged Navier-Stokes equations (RANS), supplemented
with turbulence models (such as the k-ε model), are used in practical computational fluid
dynamics (CFD) applications when modeling turbulent flows. Another technique for solving
numerically the Navier–Stokes equation is the Large-eddy simulation (LES). This approach is

17
computationally more expensive than the RANS method (in time and computer memory), but
produces better results since the larger turbulent scales are explicitly resolved.

Applicability:

Together with supplemental equations (for example, conservation of mass) and well
formulated boundary conditions, the Navier–Stokes equations seem to model fluid motion
accurately; even turbulent flows seem (on average) to agree with real world observations.

The Navier–Stokes equations assume that the fluid being studied is a continuum not moving
at relativistic velocities. At very small scales or under extreme conditions, real fluids made
out of discrete molecules will produce results different from the continuous fluids modeled
by the Navier–Stokes equations. Depending on the Knudsen number of the problem,
statistical mechanics or possibly even molecular dynamics may be a more appropriate
approach.

Another limitation is very simply the complicated nature of the equations. Time tested
formulations exist for common fluid families, but the application of the Navier–Stokes
equations to less common families tends to result in very complicated formulations which are
an area of current research. For this reason, the Navier–Stokes equations are usually written
for Newtonian fluids.

Almost universally the equations are written for a simple class of fluids (which most liquids
and all known gases belong to) known as Newtonian fluids. Studying such fluids is "simple"
because the viscosity model ends up being linear; truly general models for the flow of other
kinds of fluids (such as blood) do not, as of 2009, exist.

Derivation and description:

The derivation of the Navier–Stokes equations begins with an application of the conservation
of momentum (often alongside mass and energy conservation) being written for an arbitrary
control volume. In an inertial frame of reference, the most general form of the Navier–Stokes
equation ends up being:

where is the flow velocity, ρ is the fluid density, p is the pressure, is the (deviatoric)
stress tensor, and represents body forces (per unit volume) acting on the fluid and is the
del operator. This is a statement of the conservation of momentum in a fluid and it is an
application of Newton's second law to a continuum; in fact this equation is applicable to any
non-relativistic continuum and is known as the Cauchy momentum equation.

This equation is often written using the substantive derivative, making it more apparent that
this is a statement of Newton's law:

18
The left side of the equation describes acceleration, and may be composed of time dependent
or convective effects (also the effects of non-inertial coordinates if present). The right side of
the equation is in effect a summation of body forces (such as gravity) and divergence of stress
(pressure and stress).

Convective acceleration:

An example of convection. Though the flow is steady (time independent), the fluid decelerates as it
moves down the diverging duct (when the flow is subsonic), hence there is acceleration.

A very significant feature of the Navier–Stokes equations is the presence of convective


acceleration: the effect of time independent acceleration of a fluid with respect to space,
represented by the nonlinear quantity:

which may be interpreted either as or as with the tensor derivative


of the velocity vector Both interpretations give the same result, independent of the
coordinate system — provided is interpreted as the covariant derivative.[1]

Interpretation as (v·∇) v:

The convection term is often written as

where the advection operator is used. Usually this representation is preferred because it
[1]
is simpler than the one in terms of the tensor derivative

Interpretation as v·(∇v):

19
Here is the tensor derivative of the velocity vector, equal in Cartesian coordinates to the
component by component gradient. The convection term may, by a vector calculus identity,
be expressed without a tensor derivative:[2][3]

The form has use in irrotational flow, where the curl of the velocity (called vorticity)
is equal to zero.

Regardless of what kind of fluid is being dealt with, convective acceleration is a nonlinear
effect. Convection is present in most flows, exceptions include creeping flow (also called
Stokes flow) and incompressible flow in one dimension.

Stresses:

The effect of stress in the fluid is represented by the and terms, these are gradients of
surface forces, analogous to stresses in a solid. is called the pressure gradient and arises
from normal stresses that turn up in almost all situations, dynamic or not. conventionally
describes viscous forces; for incompressible flow, this is only a shear effect. is the
deviatoric stress tensor, and the stress tensor is equal to:

Where is the identity matrix. Interestingly, only the gradient of pressure matters, not
the pressure itself. The effect of the pressure gradient is that fluid flows from high pressure to
low pressure.

The stress derivative term contains too many unknowns to be immediately usable, so the
general form of the Navier–Stokes equation is not immediately usable to solve problems. For
this reason, assumptions on the specific viscous behavior of a fluid are made (based on
natural observations) and applied in order to specify this quantity in terms of familiar
variables, such as velocity. For example, this term becomes the useful quantity when
the fluid is assumed incompressible and Newtonian.

Other forces:

represents "other" (body force) forces. Typically this is only gravity, but may include other
fields (such as electromagnetic). In a non-inertial coordinate system, other "forces" such as
that associated with rotating coordinates may be inserted.

Often, these forces may be represented as the gradient of some scalar quantity. Gravity in the
z direction, for example, is the gradient of ρgz. Since pressure shows up only as a gradient,
this implies that solving a problem without any such body force can be amended to include
the body force by modifying pressure.

Other equations:

20
The Navier–Stokes equations are strictly a statement of the conservation of momentum. In
order to fully describe fluid flow, more information is needed (how much depends on the
assumptions made), this may include boundary data (no-slip, capillary surface, etc), the
conservation of mass, the conservation of energy, and/or an equation of state.

Regardless of the flow assumptions, a statement of the conservation of mass is generally


necessary. This is achieved through the mass continuity equation, given in its most general
form as:

Or, using the substantive derivative

Incompressible flow of Newtonian fluids:

A simplification of the resulting flow equations is obtained when considering an


incompressible flow of a Newtonian fluid. The assumption of incompressibility rules out the
possibility of sound or shock waves to occur; so this simplification is invalid if these
phenomena are important. The incompressible flow assumption typically holds well even
when dealing with a "compressible" fluid — such as air at room temperature — at low Mach
numbers (even when flowing up to about Mach 0.3). Taking the incompressible flow
assumption into account and assuming constant viscosity, the Navier–Stokes equations will
read, in vector form:[4]

Here f represents "other" body forces (forces per unit volume), such as gravity or centrifugal
force. It's well worth observing the meaning of each term (compare to the Cauchy momentum
equation):

Note that only the convective terms are nonlinear for incompressible Newtonian flow. The
convective acceleration is an acceleration caused by a (possibly steady) change in velocity
over position, for example the speeding up of fluid entering a converging nozzle. Though
individual fluid particles are being accelerated and thus are under unsteady motion, the flow
field (a velocity distribution) will not necessarily be time dependent.

21
Another important observation is that the viscosity is represented by the vector Laplacian of
the velocity field. This implies that Newtonian viscosity is diffusion of momentum, this
works in much the same way as the diffusion of heat seen in the heat equation (which also
involves the Laplacian).

If temperature effects are also neglected, the only "other" equation (apart from
initial/boundary conditions) needed is the mass continuity equation. Under the
incompressible assumption, density is a constant and it follows that the equation will simplify
to:

This is more specifically a statement of the conservation of volume (see divergence).

These equations are commonly used in 3 coordinates systems: Cartesian, cylindrical, and
spherical. While the Cartesian equations seem to follow directly from the vector equation
above, the vector form of the Navier–Stokes equation involves some tensor calculus which
means that writing it in other coordinate systems is not as simple as doing so for scalar
equations (such as the heat equation).

Cartesian coordinates:

Writing the vector equation explicitly,

Note that gravity has been accounted for as a body force, and the values of gx, gy, gz will
depend on the orientation of gravity with respect to the chosen set of coordinates.

The continuity equation reads:

The velocity components (the dependent variables to be solved for) are typically named u, v,
and w. This system of four equations comprises the most commonly used and studied form.
Though comparatively more compact than other representations, this is a nonlinear system of
partial differential equations for which solutions are difficult to obtain.

Cylindrical coordinates:
22
A change of variables on the Cartesian equations will yield the following momentum
equations for r, θ, and z:

The gravity components will generally not be constants, however for most applications either
the coordinates are chosen so that the gravity components are constant or else it is assumed
that gravity is counteracted by a pressure field (for example, flow in horizontal pipe is treated
normally without gravity and without a vertical pressure gradient). The continuity equation
is:

This cylindrical representation of the incompressible Navier–Stokes equations is the second


most commonly seen (the first being Cartesian above). Cylindrical coordinates are chosen to
take advantage of symmetry, so that a velocity component can disappear. A very common
case is axisymmetric flow, where there is no tangential velocity (uθ = 0) and the remaining
quantities are independent of θ:

Spherical coordinates:

In spherical coordinates, the r, θ, and φ momentum equations are (note the convention used:
φ is colatitude):

23
Mass continuity will read:

These equations could be (slightly) compacted by, for example, factoring 1 / r2 from the
viscous terms. This isn't done to preserve the structure of the Laplacian and other quantities.

Stream function formulation:

Taking the curl of the Navier–Stokes equation results in the elimination of pressure. This is
especially easy to see if 2D Cartesian flow is assumed (w = 0 and no dependence of anything
on z), where the equations reduce to:

Differentiating the first with respect to y, the second with respect to x and subtracting the
resulting equations will eliminate pressure and any potential force. Defining the stream
function ψ through

24
results in mass continuity being unconditionally satisfied (given the stream function is
continuous), and then incompressible Newtonian 2D momentum and mass conservation
degrade into one equation:

Where is the (2D) biharmonic operator and ν is the kinematic viscosity, This
single equation together with appropriate boundary conditions describes 2D fluid flow, taking
only kinematic viscosity as a parameter. Note that the equation for creeping flow results
when the left side is assumed zero.

In axisymmetric flow another stream function formulation, called the Stokes stream function,
can be used to describe the velocity components of an incompressible flow with one scalar
function.

Compressible flow of Newtonian fluids:

There are some exceptional phenomena that are closely linked with fluid compressibility.
One of the obvious examples is sound. Description of such phenomena requires more general
presentation of the Navier–Stokes equation that takes into account fluid compressibility. If
viscosity is assumed a constant, one additional term appears, as shown here:

where μv is second viscosity coefficient. It is related to volume viscosity or bulk viscosity.


This additional term disappears for incompressible fluid, when the divergence of the flow
equals 0.

Application to specific problems:

The Navier–Stokes equations, even when written explicitly for specific fluids, are rather
generic in nature and their proper application to specific problems can be very diverse. This is
partly because there is an enormous variety of problems that may be modeled, ranging from
as simple as the distribution of static pressure to as complicated as multiphase flow driven by
surface tension.

Generally, application to specific problems begins with some flow assumptions and
initial/boundary condition formulation, this may be followed by scale analysis to further
simplify the problem. For example, after assuming steady, parallel, one dimensional,
nonconvective pressure driven flow between parallel plates, the resulting scaled
(dimensionless) boundary value problem is:

25
Visualization of a) parallel flow and b) radial flow

The boundary condition is the no slip condition. This problem is easily solved for the flow
field:

From this point onward more quantities of interest can be easily obtained, such as viscous
drag force or net flow rate.

Difficulties may arise when the problem becomes slightly more complicated. A seemingly
modest twist on the parallel flow above would be the radial flow between parallel plates; this
involves convection and thus nonlinearity. The velocity field may be represented by a
function f(z) that must satisfy:

This ordinary differential equation is what is obtained when the Navier–Stokes equations are
written and the flow assumptions applied (additionally, the pressure gradient is solved for).
The nonlinear term makes this a very difficult problem to solve analytically (a lengthy
implicit solution may be found which involves elliptic integrals and roots of cubic
26
polynomials). Issues with the actual existence of solutions arise for R > 1.41 (approximately.
This is not the square root of two), the parameter R being the Reynolds number with
appropriately chosen scales. This is an example of flow assumptions losing their
applicability, and an example of the difficulty in "high" Reynolds number flows.

Exact solutions of the Navier–Stokes equations:

Some exact solutions to the Navier–Stokes equations exist. Examples of degenerate cases —
with the non-linear terms in the Navier–Stokes equations equal to zero — are Poiseuille flow,
Couette flow and the oscillatory Stokes boundary layer. But also more interesting examples,
solutions to the full non-linear equations, exist; for example the Taylor–Green vortex. Note
that the existence of these exact solutions does not imply they are stable: turbulence may
develop at higher Reynolds numbers.

Fluid mechanics:
Fluid mechanics is the study of how fluids move and the forces on them. (Fluids include
liquids and gases.) Fluid mechanics can be divided into fluid statics, the study of fluids at
rest, and fluid dynamics, the study of fluids in motion. It is a branch of continuum mechanics,
a subject which models matter without using the information that it is made out of atoms.
Fluid mechanics, especially fluid dynamics, is an active field of research with many unsolved
or partly solved problems. Fluid mechanics can be mathematically complex. Sometimes it
can best be solved by numerical methods, typically using computers. A modern discipline,
called Computational Fluid Dynamics (CFD), is devoted to this approach to solving fluid
mechanics problems. Also taking advantage of the highly visual nature of fluid flow is
Particle Image Velocimetry, an experimental method for visualizing and analyzing fluid flow.
Fluid mechanics is the branch of physics which deals with the properties of fluids, namely
liquids and gases, and their interaction with forces.

Brief history:

The study of fluid mechanics goes back at least to the days of ancient Greece, when
Archimedes investigated fluid statics and buoyancy. Medieval Persian natural philosophers,
including Abū Rayhān al-Bīrūnī and Al-Khazini, combined that earlier work with dynamics[1]
to presage the later development of fluid dynamics. Rapid advancement in fluid mechanics
began with Leonardo da Vinci (observation and experiment), Evangelista Torricelli
(barometer), Isaac Newton (viscosity) and Blaise Pascal (hydrostatics), and was continued by

27
Daniel Bernoulli with the introduction of mathematical fluid dynamics in Hydrodynamica
(1738). Inviscid flow was further analyzed by various mathematicians (Leonhard Euler,
d'Alembert, Lagrange, Laplace, Poisson) and viscous flow was explored by a multitude of
engineers including Poiseuille and Gotthilf Heinrich Ludwig Hagen. Further mathematical
justification was provided by Claude-Louis Navier and George Gabriel Stokes in the Navier-
Stokes Equations, and boundary layers were investigated (Ludwig Prandtl), while various
scientists (Osborne Reynolds, Andrey Kolmogorov, Geoffrey Ingram Taylor) advanced the
understanding of fluid viscosity and turbulence.

Relationship to continuum mechanics:

Fluid mechanics is a subdiscipline of continuum mechanics, as illustrated in the following


table.

Elasticity: which describes materials that return to


their rest shape after an applied stress
Solid mechanics: the study
of the physics of
Plasticity: which describes
continuous materials with a
materials that permanently
defined rest shape.
Continuum deform after a large Rheology: the study of
mechanics the study enough force is applied materials with both
of the physics of stress. solid and fluid
continuous materials characteristics

Fluid mechanics: the Non-Newtonian fluids


study of the physics of
continuous materials
which take the shape of
Newtonian fluids
their container.

In a mechanical view, a fluid is a substance that does not support tangential stress; that is why
a fluid at rest has the shape of its containing vessel. A fluid at rest has no shear stress.

28
Assumptions:

Like any mathematical model of the real world, fluid mechanics makes some basic
assumptions about the materials being studied. These assumptions are turned into equations
that must be satisfied if the assumptions are to hold true. For example, consider an
incompressible fluid in three dimensions. The assumption that mass is conserved means that
for any fixed closed surface (such as a sphere) the rate of mass passing from outside to inside
the surface must be the same as rate of mass passing the other way. (Alternatively, the mass
inside remains constant, as does the mass outside). This can be turned into an integral
equation over the surface.

Fluid mechanics assumes that every fluid obeys the following:

• Conservation of mass
• Conservation of momentum
• The continuum hypothesis, detailed below.

Further, it is often useful (and realistic) to assume a fluid is incompressible - that is, the
density of the fluid does not change. Liquids can often be modelled as incompressible fluids,
whereas gases cannot.

Similarly, it can sometimes be assumed that the viscosity of the fluid is zero (the fluid is
inviscid). Gases can often be assumed to be inviscid. If a fluid is viscous, and its flow
contained in some way (e.g. in a pipe), then the flow at the boundary must have zero velocity.
For a viscous fluid, if the boundary is not porous, the shear forces between the fluid and the
boundary results also in a zero velocity for the fluid at the boundary. This is called the no-slip
condition. For a porous media otherwise, in the frontier of the containing vessel, the slip
condition is not zero velocity, and the fluid has a discontinuous velocity field between the
free fluid and the fluid in the porous media (this is related to the Beavers and Joseph
condition).

The continuum hypothesis:

Fluids are composed of molecules that collide with one another and solid objects. The
continuum assumption, however, considers fluids to be continuous. That is, properties such as
density, pressure, temperature, and velocity are taken to be well-defined at "infinitely" small
points, defining a REV (Reference Element of Volume), at the geometric order of the
distance between two adjacent molecules of fluid. Properties are assumed to vary
continuously from one point to another, and are averaged values in the REV. The fact that the
fluid is made up of discrete molecules is ignored.

The continuum hypothesis is basically an approximation, in the same way planets are
approximated by point particles when dealing with celestial mechanics, and therefore results
in approximate solutions. Consequently, assumption of the continuum hypothesis can lead to
results which are not of desired accuracy. That said, under the right circumstances, the
continuum hypothesis produces extremely accurate results.

Those problems for which the continuum hypothesis does not allow solutions of desired
accuracy are solved using statistical mechanics. To determine whether or not to use

29
conventional fluid dynamics or statistical mechanics, the Knudsen number is evaluated for
the problem. The Knudsen number is defined as the ratio of the molecular mean free path
length to a certain representative physical length scale. This length scale could be, for
example, the radius of a body in a fluid. (More simply, the Knudsen number is how many
times its own diameter a particle will travel on average before hitting another particle).
Problems with Knudsen numbers at or above unity are best evaluated using statistical
mechanics for reliable solutions.

Navier-Stokes equations:

The Navier-Stokes equations (named after Claude-Louis Navier and George Gabriel Stokes)
are the set of equations that describe the motion of fluid substances such as liquids and gases.
These equations state that changes in momentum (force) of fluid particles depend only on the
external pressure and internal viscous forces (similar to friction) acting on the fluid. Thus, the
Navier-Stokes equations describe the balance of forces acting at any given region of the fluid.

The Navier-Stokes equations are differential equations which describe the motion of a fluid.
Such equations establish relations among the rates of change the variables of interest. For
example, the Navier-Stokes equation for an ideal fluid with zero viscosity states that
acceleration (the rate of change of velocity) is proportional to the derivative of internal
pressure.

This means that solutions of the Navier-Stokes equations for a given physical problem must
be sought with the help of calculus. In practical terms only the simplest cases can be solved
exactly in this way. These cases generally involve non-turbulent, steady flow (flow does not
change with time) in which the Reynolds number is small.

For more complex situations, such as global weather systems like El Niño or lift in a wing,
solutions of the Navier-Stokes equations can currently only be found with the help of
computers. This is a field of sciences by its own called computational fluid dynamics.

General form of the equation:

The general form of the Navier-Stokes equations for the conservation of momentum is:

Where

• is the fluid density,

• is the substantive derivative (also called the material derivative),


• is the velocity vector,
• is the body force vector, and
• is a tensor that represents the surface forces applied on a fluid particle (the
comoving stress tensor).

30
Unless the fluid is made up of spinning degrees of freedom like vortices, is a symmetric
tensor. In general, (in three dimensions) has the form:

Where

• are normal stresses, and


• are tangential stresses (shear stresses).

The above is actually a set of three equations, one per dimension. By themselves, these aren't
sufficient to produce a solution. However, adding conservation of mass and appropriate
boundary conditions to the system of equations produces a solvable set of equations.

Newtonian vs. non-Newtonian fluids:

A Newtonian fluid (named after Isaac Newton) is defined to be a fluid whose shear stress is
linearly proportional to the velocity gradient in the direction perpendicular to the plane of
shear. This definition means regardless of the forces acting on a fluid, it continues to flow.
For example, water is a Newtonian fluid, because it continues to display fluid properties no
matter how much it is stirred or mixed. A slightly less rigorous definition is that the drag of a
small object being moved slowly through the fluid is proportional to the force applied to the
object. (Compare friction). Important fluids, like water as well as most gases, behave — to
good approximation — as a Newtonian fluid under normal conditions on Earth.

By contrast, stirring a non-Newtonian fluid can leave a "hole" behind. This will gradually
fill up over time - this behaviour is seen in materials such as pudding, oobleck, or sand
(although sand isn't strictly a fluid). Alternatively, stirring a non-Newtonian fluid can cause
the viscosity to decrease, so the fluid appears "thinner" (this is seen in non-drip paints). There
are many types of non-Newtonian fluids, as they are defined to be something that fails to
obey a particular property — for example, most fluids with long molecular chains can react in
a non-Newtonian manner.

Equations for a Newtonian fluid:

The constant of proportionality between the shear stress and the velocity gradient is known as
the viscosity. A simple equation to describe Newtonian fluid behaviour is

Where

τ is the shear stress exerted by the fluid ("drag")

μ is the fluid viscosity - a constant of proportionality

31
is the velocity gradient perpendicular to the direction of shear

For a Newtonian fluid, the viscosity, by definition, depends only on temperature and
pressure, not on the forces acting upon it. If the fluid is incompressible and viscosity is
constant across the fluid, the equation governing the shear stress (in Cartesian coordinates) is

Where

τij is the shear stress on the ith face of a fluid element in the jth direction

vi is the velocity in the ith direction

xj is the jth direction coordinate

If a fluid does not obey this relation, it is termed a non-Newtonian fluid, of which there are
several types.

Among fluids, two rough broad divisions can be made- ideal and non- ideal fluids. An ideal
fluid really does not exist, but in some calculations, the assumption is justifiable. An Ideal
fluid is non viscous- offers no resistance whatsoever to a shearing force.

One can group real fluids into Newtonian and non- Newtonian. Newtonian fluids agree with
Newton's law of viscosity. Non- Newtonian Fluids could either be plastic, Bingham plastic,
pseudoplastic, dilatant, thexotropic, rheopectic, viscoelatic

Hydraulics

32
Hydraulics is a topic in applied science and engineering dealing with the mechanical
properties of liquids. Fluid mechanics provides the theoretical foundation for hydraulics,
which focuses on the engineering uses of fluid properties. In fluid power, hydraulics is used
for the generation, control, and transmission of power by the use of pressurized liquids.
Hydraulic topics range through most science and engineering disciplines, and cover concepts
such as pipe flow, dam design, fluidics and fluid control circuitry, pumps, turbines,
hydropower, computational fluid dynamics, flow measurement, river channel behavior and
erosion.

Open channel hydraulics is the branch of hydraulics dealing with free surface flow, such as
occurring in rivers, canals, lakes, estuaries and seas.

The word "hydraulics" originates from the Greek word ὑδραυλικός (hydraulikos) which in
turn originates from ὕδραυλος (hydraulos) meaning water organ which in turn comes from
ὕδωρ (hydor, Greek for water) and αὐλός (aulos, meaning pipe).

Stokes flow:

33
An object moving through a gas or liquid experiences a force in direction opposite to its motion.
Terminal velocity is achieved when the drag force is equal in magnitude but opposite in direction to
the force propelling the object. Shown is a sphere in Stokes flow, at very low Reynolds number.

Stokes flow (named after George Gabriel Stokes) is a type of fluid flow where advective
inertial forces are small compared with viscous forces. The Reynolds number is low, i.e.
. This is a typical situation in flows where the fluid velocities are very slow, the
viscosities are very large, or the length-scales of the flow are very small, such as in MEMS
devices or in the flow of viscous polymers.

Stokes equations:

For this type of flow, the inertial forces are assumed to be negligible and the Navier–Stokes
equations simplify to give the Stokes equations:

where is the comoving stress tensor, and an applied body force. There is also an equation
for conservation of mass. In the common case of an incompressible Newtonian fluid, the
Stokes equations are:

Properties:

The Stokes equations represent a considerable simplification of the full Navier–Stokes


equations, especially in the incompressible Newtonian case.

34
Instantaneity

A Stokes flow has no dependence on time other than through time-dependent


boundary conditions. This means that, given the boundary conditions of a Stokes
flow, the flow can be found without knowledge of the flow at any other time.

Time-reversibility

An immediate consequence of instantaneity, time-reversibility means then a time-


reversed Stokes flow solves the same equations as the original Stokes flow. This
property can sometimes be used (in conjunction with linearity and symmetry in the
boundary conditions) to derive results about a flow without solving it fully.

While these properties are true for incompressible Newtonian Stokes flows, the non-linear
and sometimes time-dependent nature of non-Newtonian fluids means that they do not hold
in the more general case.

Methods of solution:

By stream function:

It can be shown that in 2-D, the stream function for an incompressible Newtonian Stokes
flow satisfies the biharmonic equation .

In the 3-D axisymmetric case, the Stokes stream function Ψ solves the equation E2Ψ = 0,

where

By Papkovich-Neuber solution:

The Papkovich-Neuber Solution represents the velocity and pressure fields of an


incompressible Newtonian Stokes flow in terms of two harmonic potentials.

By Boundary element method:

Certain problems, such as the evolution of the shape of a bubble in a Stokes flow, are
conducive to numerical solution by the Boundary Element method. This technique can be
applied in both 2- and 3-dimensional flows.

By Green's function:

The linearity of the Stokes equations in the case of an incompressible Newtonian fluid means
that a Green's function for the equations can be found. The solution for the pressure p and
velocity due to a point force acting at the origin with as is
given by

35
Where

is a second-rank tensor known as the Oseen Tensor (after Carl Wilhelm Oseen).

The solution for a distributed force density (again with decay at infinity) can then be
constructed by superposition:

Fluid dynamics:

Typical aerodynamic teardrop shape, showing the pressure distribution as the thickness of the black
line and showing the velocity in the boundary layer as the violet triangles. The green vortex
generators prompt the transition to turbulent flow and prevent back-flow also called flow separation
from the high pressure region in the back. The surface in front is as smooth as possible or even
employs shark like skin, as any turbulence here will reduce the energy of the airflow. The Kammback
also prevents back flow from the high pressure region in the back across the spoilers to the convergent
part. Putting stuff inside out results in tubes; they also face the problem of flow separation in their
divergent parts, so called diffusers. Cutting the shape into halves results in an aerofoil with the low
pressure region on top leading to lift (force).

In physics, fluid dynamics is the sub-discipline of fluid mechanics dealing with fluid flow
— the natural science of fluids (liquids and gases) in motion. It has several subdisciplines
itself, including aerodynamics (the study of gases in motion) and hydrodynamics (the study
of liquids in motion). Fluid dynamics has a wide range of applications, including calculating
forces and moments on aircraft, determining the mass flow rate of petroleum through
pipelines, predicting weather patterns, understanding nebulae in interstellar space and
reportedly modeling fission weapon detonation. Some of its principles are even used in traffic
engineering, where traffic is treated as a continuous fluid.
36
Fluid dynamics offers a systematic structure that underlies these practical disciplines and that
embraces empirical and semi-empirical laws, derived from flow measurement, used to solve
practical problems. The solution of a fluid dynamics problem typically involves calculation
of various properties of the fluid, such as velocity, pressure, density, and temperature, as
functions of space and time.

In an historical context, hydrodynamics had a different meaning than it has nowadays. Before
the twentieth century, hydrodynamics was synonymous with fluid dynamics. This is still
reflected in the names of some fluid dynamics topics, like magnetohydrodynamics and
hydrodynamic stability — both also applicable in, as well as being applied to, gases.

Equations of fluid dynamics:

The foundational axioms of fluid dynamics are the conservation laws, specifically,
conservation of mass, conservation of linear momentum (also known as Newton's Second
Law of Motion), and conservation of energy (also known as First Law of Thermodynamics).
These are based on classical mechanics and are modified in quantum mechanics and general
relativity. They are expressed using the Reynolds Transport Theorem.

In addition to the above, fluids are assumed to obey the continuum assumption. Fluids are
composed of molecules that collide with one another and solid objects. However, the
continuum assumption considers fluids to be continuous, rather than discrete. Consequently,
properties such as density, pressure, temperature, and velocity are taken to be well-defined at
infinitesimally small points, and are assumed to vary continuously from one point to another.
The fact that the fluid is made up of discrete molecules is ignored.

For fluids which are sufficiently dense to be a continuum, do not contain ionized species, and
have velocities small in relation to the speed of light, the momentum equations for Newtonian
fluids are the Navier-Stokes equations, which is a non-linear set of differential equations that
describes the flow of a fluid whose stress depends linearly on velocity gradients and pressure.
The unsimplified equations do not have a general closed-form solution, so they are primarily
of use in Computational Fluid Dynamics. The equations can be simplified in a number of
ways, all of which make them easier to solve. Some of them allow appropriate fluid dynamics
problems to be solved in closed form.

In addition to the mass, momentum, and energy conservation equations, a thermodynamical


equation of state giving the pressure as a function of other thermodynamic variables for the
fluid is required to completely specify the problem. An example of this would be the perfect
gas equation of state:

Where p is pressure, ρ is density, Ru is the gas constant, M is the molecular mass and T is
temperature.

Compressible vs incompressible flow:


37
All fluids are compressible to some extent that is changes in pressure or temperature will
result in changes in density. However, in many situations the changes in pressure and
temperature are sufficiently small that the changes in density are negligible. In this case the
flow can be modeled as an incompressible flow. Otherwise the more general compressible
flow equations must be used.

Mathematically, incompressibility is expressed by saying that the density ρ of a fluid parcel


does not change as it moves in the flow field, i.e.,

Where D / Dt is the substantial derivative, which is the sum of local and convective
derivatives. This additional constraint simplifies the governing equations, especially in the
case when the fluid has a uniform density.

For flow of gases, to determine whether to use compressible or incompressible fluid


dynamics, the Mach number of the flow is to be evaluated. As a rough guide, compressible
effects can be ignored at Mach numbers below approximately 0.3. For liquids, whether the
incompressible assumption is valid depends on the fluid properties (specifically the critical
pressure and temperature of the fluid) and the flow conditions (how close to the critical
pressure the actual flow pressure becomes). Acoustic problems always require allowing
compressibility, since sound waves are compression waves involving changes in pressure and
density of the medium through which they propagate.

Viscous vs inviscid flow:

Viscous problems are those in which fluid friction has significant effects on the fluid motion.

The Reynolds number can be used to evaluate whether viscous or inviscid equations are
appropriate to the problem.

Stokes flow is flow at very low Reynolds numbers, such that inertial forces can be neglected
compared to viscous forces.

On the contrary, high Reynolds numbers indicate that the inertial forces are more significant
than the viscous (friction) forces. Therefore, we may assume the flow to be an inviscid flow,
an approximation in which we neglect viscosity at all, compared to inertial terms.

This idea can work fairly well when the Reynolds number is high. However, certain problems
such as those involving solid boundaries may require that the viscosity be included. Viscosity
often cannot be neglected near solid boundaries because the no-slip condition can generate a
thin region of large strain rate (known as Boundary layer) which enhances the effect of even a
small amount of viscosity, and thus generating vorticity. Therefore, to calculate net forces on
bodies (such as wings) we should use viscous flow equations. As illustrated by d'Alembert's
paradox, a body in an inviscid fluid will experience no drag force. The standard equations of
inviscid flow are the Euler equations. Another often used model, especially in computational
fluid dynamics, is to use the Euler equations away from the body and the boundary layer
equations, which incorporates viscosity, in a region close to the body.

38
The Euler equations can be integrated along a streamline to get Bernoulli's equation. When
the flow is everywhere irrotational and inviscid, Bernoulli's equation can be used throughout
the flow field. Such flows are called potential flows.

Steady vs unsteady flow:

Hydrodynamics simulation of the Rayleigh–Taylor instability

When all the time derivatives of a flow field vanish, the flow is considered to be a steady
flow. Otherwise, it is called unsteady. Whether a particular flow is steady or unsteady can
depend on the chosen frame of reference. For instance, laminar flow over a sphere is steady
in the frame of reference that is stationary with respect to the sphere. In a frame of reference
that is stationary than the governing equations of the same problem without taking advantage
of the steadiness of the flow field.

Although strictly unsteady flows, time-periodic problems can often be solved by the same
techniques as steady flows. For this reason, they can be considered to be somewhere between
steady and unsteady.

Laminar versus turbulent flow:

Turbulence is flow characterized by recirculation, eddies, and apparent randomness. Flow in


which turbulence is not exhibited is called laminar. It should be noted, however, that the
presence of eddies or recirculation alone does not necessarily indicate turbulent flow—these
phenomena may be present in laminar flow as well. Mathematically, turbulent flow is often
represented via a Reynolds decomposition, in which the flow is broken down into the sum of
an average component and a perturbation component.

It is believed that turbulent flows can be described well through the use of the Navier–Stokes
equations. Direct numerical simulation (DNS), based on the Navier–Stokes equations makes
it possible to simulate turbulent flows at moderate Reynolds numbers. Restrictions depend on
the power of the computer used and the efficiency of the solution algorithm. The results of
DNS agree with the experimental data.

39
Most flows of interest have Reynolds numbers much too high for DNS to be a viable
option[3], given the state of computational power for the next few decades. Any flight vehicle
large enough to carry a human (L > 3 m), moving faster than 72 km/h (20 m/s) is well beyond
the limit of DNS simulation (Re = 4 million). Transport aircraft wings (such as on an Airbus
A300 or Boeing 747) have Reynolds numbers of 40 million (based on the wing chord). In
order to solve these real-life flow problems, turbulence models will be a necessity for the
foreseeable future. Reynolds-averaged Navier–Stokes equations (RANS) combined with
turbulence modeling provides a model of the effects of the turbulent flow. Such a modeling
mainly provides the additional momentum transfer by the Reynolds stresses, although the
turbulence also enhances the heat and mass transfer. Another promising methodology is large
eddy simulation (LES), especially in the guise of detached eddy simulation (DES)—which is
a combination of RANS turbulence modeling and large eddy simulation.

Newtonian vs non-Newtonian fluids:

Sir Isaac Newton showed how stress and the rate of strain are very close to linearly related
for many familiar fluids, such as water and air. These Newtonian fluids are modeled by a
coefficient called viscosity, which depends on the specific fluid.

However, some of the other materials, such as emulsions and slurries and some visco-elastic
materials (e.g. blood, some polymers), have more complicated non-Newtonian stress-strain
behaviours. These materials include sticky liquids such as latex, honey, and lubricants which
are studied in the sub-discipline of rheology.

Subsonic vs transonic, supersonic and hypersonic flows:

While many terrestrial flows (e.g. flow of water through a pipe) occur at low mach numbers,
many flows of practical interest (e.g. in aerodynamics) occur at high fractions of the Mach
Number M=1 or in excess of it (supersonic flows). New phenomena occur at these Mach
number regimes (e.g. shock waves for supersonic flow, transonic instability in a regime of
flows with M nearly equal to 1, non-equilibrium chemical behavior due to ionization in
hypersonic flows) and it is necessary to treat each of these flow regimes separately.

Non-relativistic vs relativistic flows:

Classical fluid dynamics is derived based on Newtonian mechanics, which is adequate for
most applications. However, at speeds comparable to the speed of light, c, Newtonian
mechanics is inaccurate and a relativistic framework has to be used instead.

Magnetohydrodynamics:

Magnetohydrodynamics is the multi-disciplinary study of the flow of electrically conducting


fluids in electromagnetic fields. Examples of such fluids include plasmas, liquid metals, and
salt water. The fluid flow equations are solved simultaneously with Maxwell's equations of
electromagnetism.

Other approximations:

40
There are a large number of other possible approximations to fluid dynamic problems. Some
of the more commonly used are listed below.

• The Boussinesq approximation neglects variations in density except to calculate


buoyancy forces. It is often used in free convection problems where density changes
are small.
• Lubrication theory and Hele-Shaw flow exploits the large aspect ratio of the domain
to show that certain terms in the equations are small and so can be neglected.
• Slender-body theory is a methodology used in Stokes flow problems to estimate the
force on, or flow field around, a long slender object in a viscous fluid.
• The shallow-water equations can be used to describe a layer of relatively inviscid
fluid with a free surface, in which surface gradients are small.
• The Boussinesq equations are applicable to surface waves on thicker layers of fluid
and with steeper surface slopes.
• Darcy's law is used for flow in porous media, and works with variables averaged
over several pore-widths.
• In rotating systems, the quasi-geostrophic approximation assumes an almost perfect
balance between pressure gradients and the Coriolis force. It is useful in the study of
atmospheric dynamics.

Terminology in fluid dynamics:

The concept of pressure is central to the study of both fluid statics and fluid dynamics. A
pressure can be identified for every point in a body of fluid, regardless of whether the fluid is
in motion or not. Pressure can be measured using an aneroid, Bourdon tube, mercury column,
or various other methods.

Some of the terminology that is necessary in the study of fluid dynamics is not found in other
similar areas of study. In particular, some of the terminology used in fluid dynamics is not
used in fluid statics.

Terminology in incompressible fluid dynamics:

The concepts of total pressure and dynamic pressure arise from Bernoulli's equation and are
significant in the study of all fluid flows. (These two pressures are not pressures in the usual
sense - they cannot be measured using an aneroid, Bourdon tube or mercury column.) To
avoid potential ambiguity when referring to pressure in fluid dynamics, many authors use the
term static pressure to distinguish it from total pressure and dynamic pressure. Static pressure
is identical to pressure and can be identified for every point in a fluid flow field.

In Aerodynamics, L.J. Clancy writes: To distinguish it from the total and dynamic pressures,
the actual pressure of the fluid, which is associated not with its motion but with its state, is
often referred to as the static pressure, but where the term pressure alone is used it refers to
this static pressure.

A point in a fluid flow where the flow has come to rest (i.e. speed is equal to zero adjacent to
some solid body immersed in the fluid flow) is of special significance. It is of such
41
importance that it is given a special name - a stagnation point. The static pressure at the
stagnation point is of special significance and is given its own name - stagnation pressure. In
incompressible flows, the stagnation pressure at a stagnation point is equal to the total
pressure throughout the flow field.

Terminology in compressible fluid dynamics:

In a compressible fluid, such as air, the temperature and density are essential when
determining the state of the fluid. In addition to the concept of total pressure (also known as
stagnation pressure), the concepts of total (or stagnation) temperature and total (or stagnation)
density are also essential in any study of compressible fluid flows. To avoid potential
ambiguity when referring to temperature and density, many authors use the terms static
temperature and static density. Static temperature is identical to temperature; and static
density is identical to density; and both can be identified for every point in a fluid flow field.

The temperature and density at a stagnation point are called stagnation temperature and
stagnation density.

A similar approach is also taken with the thermodynamic properties of compressible fluids.
Many authors use the terms total (or stagnation) enthalpy and total (or stagnation) entropy.
The terms static enthalpy and static entropy appear to be less common, but where they are
used they mean nothing more than enthalpy and entropy respectively, and the prefix 'static' is
being used to avoid ambiguity with their 'total' or 'stagnation' counterparts.

Fluid:
A fluid is defined as a substance that continually deforms (flows) under an applied shear
stress. All gases are fluids, but not all liquids are fluids. Fluids are a subset of the phases of
matter and include liquids, gases, plasmas and, to some extent, plastic solids.

In common usage, "fluid" is often used as a synonym for "liquid", with no implication that
gas could also be present. For example, "brake fluid" is hydraulic oil and will not perform its
required function if there is gas in it. This colloquial usage of the term is also common in
medicine ("take plenty of fluids"), and in nutrition.

Liquids form a free surface (that is, a surface not created by the container) while gases do not.
The distinction between solids and fluid is not entirely obvious. The distinction is made by
evaluating the viscosity of the substance. Silly Putty can be considered to behave like a solid
or a fluid, depending on the time period over which it is observed. It is best described as a
viscoelastic fluid.

42
Fluids display such properties as:

• not resisting deformation, or resisting it only lightly (viscosity), and


• The ability to flow (also described as the ability to take on the shape of the container).

These properties are typically a function of their inability to support a shear stress in static
equilibrium.

Solids can be subjected to shear stresses, and to normal stresses - both compressive and
tensile. In contrast, ideal fluids can only be subjected to normal, compressive stress which is
called pressure. Real fluids display viscosity and so are capable of being subjected to low
levels of shear stress.

In a solid, shear stress is a function of strain, but in a fluid, shear stress is a function of rate of
strain. A consequence of this behavior is Pascal's law which describes the role of pressure in
characterizing a fluid's state.

Depending on the relationship between shear stress, and the rate of strain and its derivatives,
fluids can be characterized as:

• Newtonian fluids : where stress is directly proportional to rate of strain, and


• Non-Newtonian fluids : where stress is proportional to rate of strain, its higher
powers and derivatives.

The behavior of fluids can be described by the Navier-Stokes equations—a set of partial
differential equations which are based on:

• continuity (conservation of mass),


• conservation of linear momentum
• conservation of angular momentum
• Conservation of energy.

The study of fluids is fluid mechanics, which is subdivided into fluid dynamics and fluid
statics depending on whether the fluid is in motion.

Turbulence:

43
Turbulent flow around an obstacle; the flow farther upstream is laminar

Laminar and turbulent water flow over the hull of a submarine

Turbulence in the tip vortex from an airplane wing

In fluid dynamics, turbulence or turbulent flow is a fluid regime characterized by chaotic,


stochastic property changes. This includes low momentum diffusion, high momentum
convection, and rapid variation of pressure and velocity in space and time. Flow that is not
turbulent is called laminar flow. The (dimensionless) Reynolds number characterizes whether
flow conditions lead to laminar or turbulent flow; e.g. for pipe flow, a Reynolds number
above about 4000 (A Reynolds number between 2100 and 4000 is known as transitional
flow) will be turbulent. At very low speeds the flow is laminar, i.e., the flow is smooth
(though it may involve vortices on a large scale). As the speed increases, at some point the
transition is made to turbulent flow. In turbulent flow, unsteady vortices appear on many
scales and interact with each other. Drag due to boundary layer skin friction increases. The
44
structure and location of boundary layer separation often changes, sometimes resulting in a
reduction of overall drag. Because laminar-turbulent transition is governed by Reynolds
number, the same transition occurs if the size of the object is gradually increased, or the
viscosity of the fluid is decreased, or if the density of the fluid is increased.

In natural fluids such as the ocean, atmosphere and in stars, inertial vortex forces vxw that
drive turbulence are restricted by gravity and rotation at large scales to form fossil turbulence.
To take this into account, a narrow definition of turbulence is needed.

Turbulence is defined as an eddy-like state of fluid motion where the inertial vortex forces
of the eddies are larger than any of the other forces that tend to damp the eddies out. By this
definition, irrotational flows are non-turbulent. The energy cascade from large scales to small
in the irrotational flow is not a turbulence cascade. Thus, turbulence always cascades from
small scales to large; that is, from the Kolmogorov scale where vorticity is produced to the
Oboukov or energy scale, which monotonically grows with time. In stably stratified flows,
buoyancy forces limit the vertical overturn scales of turbulence at the Ozmidov scale where
the turbulent kinetic energy is converted to saturated internal wave motions termed fossil
vorticity turbulence. The first turbulence and first turbulent combustion produced the big
bang.

Converted oceanic and atmospheric turbulent kinetic energy is radiated at about 45 degree
angles as nonlinear fossil turbulence internal waves with the Ozmidov scale at fossilization.
The physical mechanism for the turbulence cascade from small scales to large is merging of
adjacent eddies with the same spin, driven together by inertial vortex forces. Fossil
turbulence is defined as a perturbation in any hydrophysical field caused by turbulence that
persists after the fluid is no longer turbulent at the scale of the perturbation. Inflation beyond
the scale of causal connection ct caused the first fossil turbulence, where c is the speed of
light and t is the time after the big bang.

Turbulence (obsolete definition) causes the formation of eddies of many different length
scales. Most of the kinetic energy of the turbulent motion is contained in the large scale
structures. The energy "cascades" from these large scale structures to smaller scale structures
by an inertial and essentially inviscid mechanism (obsolete). This process continues, creating
smaller and smaller structures which produces a hierarchy of eddies. Eventually this process
creates structures that are small enough that molecular diffusion becomes important and
viscous dissipation of energy finally takes place. The scale at which this happens is the
Kolmogorov length scale.

In two dimensional turbulence (as can be approximated in the atmosphere or ocean), energy
actually flows to larger scales. This is referred to as the inverse energy cascade (obsolete
terminology) and is characterized by a k − (5 / 3) in the power spectrum. This is the main reason
why large scale weather features such as hurricanes occur.

Turbulent diffusion is usually described by a turbulent diffusion coefficient. This turbulent


diffusion coefficient is defined in a phenomenological sense, by analogy with the molecular
diffusivities, but it does not have a true physical meaning, being dependent on the flow
conditions, and not a property of the fluid, itself. In addition, the turbulent diffusivity concept
assumes a constitutive relation between a turbulent flux and the gradient of a mean variable
similar to the relation between flux and gradient that exists for molecular transport. In the
best case, this assumption is only an approximation. Nevertheless, the turbulent diffusivity is
45
the simplest approach for quantitative analysis of turbulent flows, and many models have
been postulated to calculate it. For instance, in large bodies of water like oceans this
coefficient can be found using Richardson's four-third power law and is governed by the
random walk principle. In rivers and large ocean currents, the diffusion coefficient is given
by variations of Elder's formula.

When designing piping systems, turbulent flow requires a higher input of energy from a
pump (or fan) than laminar flow. However, for applications such as heat exchangers and
reaction vessels, turbulent flow is essential for good heat transfer and mixing.

While it is possible to find some particular solutions of the Navier-Stokes equations


governing fluid motion, all such solutions are unstable at large Reynolds numbers. Sensitive
dependence on the initial and boundary conditions makes fluid flow irregular both in time
and in space so that a statistical description is needed. Russian mathematician Andrey
Kolmogorov proposed the first statistical theory of turbulence, based on the aforementioned
notion of the energy cascade (an idea originally introduced by Richardson) and the concept of
self-similarity. As a result, the Kolmogorov microscales were named after him. It is now
known that the self-similarity is broken so the statistical description is presently modified.
Still, the complete description of turbulence remains one of the unsolved problems in physics.
According to an apocryphal story Werner Heisenberg was asked what he would ask God,
given the opportunity. His reply was: "When I meet God, I am going to ask him two
questions: Why relativity? And why turbulence? I really believe he will have an answer for
the first." A similar witticism has been attributed to Horace Lamb (who had published a noted
text book on Hydrodynamics) — his choice being quantum electrodynamics (instead of
relativity) and turbulence. Lamb was quoted as saying in a speech to the British Association
for the Advancement of Science, "I am an old man now, and when I die and go to heaven
there are two matters on which I hope for enlightenment. One is quantum electrodynamics,
and the other is the turbulent motion of fluids. And about the former I am rather optimistic."
The driving force for turbulence (modern definition) is the Lamb vector vxw.

Turbulent mixing and turbulent diffusivity

The most important aspect of turbulent flows is their ability to rapidly mix and diffuse scalar
fluid properties such as temperature and chemical species. Universal similarity laws for
turbulent mixing have been proposed and experimentally tested which are similar to those of
Kolmogorov-Obukhov 1941 and 1962.

46
Examples of turbulence:

Laminar and turbulent flow of cigarette smoke.

• Smoke rising from a cigarette. For the first few centimeters, the flow remains laminar,
and then becomes unstable and turbulent as the rising hot air accelerates upwards.
Similarly, the dispersion of pollutants in the atmosphere is governed by turbulent
processes.

• Flow over a golf ball. (This can be best understood by considering the golf ball to be
stationary, with air flowing over it.) If the golf ball were smooth, the boundary layer
flow over the front of the sphere would be laminar at typical conditions. However, the
boundary layer would separate early, as the pressure gradient switched from favorable
(pressure decreasing in the flow direction) to unfavorable (pressure increasing in the
flow direction), creating a large region of low pressure behind the ball that creates
high form drag. To prevent this from happening, the surface is dimpled to perturb the
boundary layer and promote transition to turbulence. This results in higher skin
friction, but moves the point of boundary layer separation further along, resulting in
lower form drag and lower overall drag.

• The mixing of warm and cold air in the atmosphere by wind, which causes clear-air
turbulence experienced during airplane flight, as well as poor astronomical seeing (the
blurring of images seen through the atmosphere.)

• Most of the terrestrial atmospheric circulation


• The oceanic and atmospheric mixed layers and intense oceanic currents.
• The flow conditions in many industrial equipment (such as pipes, ducts, precipitators,
gas scrubbers, dynamic scraped surface heat exchangers, etc.) and machines (for
instance, internal combustion engines and gas turbines).

47
• The external flow over all kind of vehicles such as cars, airplanes, ships and
submarines.
• The motions of matter in stellar atmospheres.
• Spinning motions of Planck particles and anti-particles during big bang turbulent
combustion.
• A jet exhausting from a nozzle into a quiescent fluid. As the flow emerges into this
external fluid, shear layers originating at the lips of the nozzle are created. These
layers separate the fast moving jet from the external fluid, and at a certain critical
Reynolds number they become unstable and break down to turbulence.

Unsolved problems in physics: Is it possible to make a theoretical model to describe the


behavior of a turbulent flow — in particular, its internal structures?

• Race cars unable to follow each other through fast corners due to turbulence created
by the leading car causing understeer.
• In windy conditions, trucks that are on the motorway gets buffeted by their wake.
• Round bridge supports under water. In the summer when the river is flowing slowly
the water goes smoothly around the support legs. In the winter the flow is faster, so a
higher Reynolds Number, so the flow may start off laminar but is quickly separated
from the leg and becomes turbulent.

Kolmogorov 1941 Theory:

Richardson's notion of turbulence was that a turbulent flow is composed by "eddies" of


different sizes. The sizes define a characteristic length scale for the eddies, which are also
characterized by velocity scales and time scales (turnover time) dependent on the length
scale. The large eddies are unstable and eventually break up originating smaller eddies, and
the kinetic energy of the initial large eddy is divided into the smaller eddies that stemmed
from it. These smaller eddies undergo the same process, giving rise to even smaller eddies
which inherit the energy of their predecessor eddy, and so on. In this way, the energy is
passed down from the large scales of the motion to smaller scales until reaching a sufficiently
small length scale such that the viscosity of the fluid can effectively dissipate the kinetic
energy into internal energy.

In his original theory of 1941, Kolmogorov postulated that for very high Reynolds number,
the small scale turbulent motions are statistically isotropic (i.e. no preferential spatial
direction could be discerned). In general, the large scales of a flow are not isotropic, since
they are determined by the particular geometrical features of the boundaries (the size
characterizing the large scales will be denoted as L). Kolmogorov's idea was that in the
Richardson's energy cascade this geometrical and directional information is lost, while the
scale is reduced, so that the statistics of the small scales has a universal character: they are the
same for all turbulent flows when the Reynolds number is sufficiently high.

Thus, Kolmogorov introduced a second hypothesis: for very high Reynolds numbers the
statistics of small scales are universally and uniquely determined by the viscosity (ν) and the

48
rate of energy dissipation ( ). With only these two parameters, the unique length that can be
formed by dimensional analysis is

This is today known as the Kolmogorov length scale (see Kolmogorov microscales).

A turbulent flow (obsolete definition) is characterized by a hierarchy of scales through which


the energy cascade takes place. Dissipation of kinetic energy takes place at scales of the order
of Kolmogorov length η, while the input of energy into the cascade comes from the decay of
the large scales, of order L. These two scales at the extremes of the cascade can differ by
several orders of magnitude at high Reynolds numbers. In between there is a range of scales
(each one with its own characteristic length r) that has formed at the expense of the energy of
the large ones. These scales are very large compared with the Kolmogorov length, but still
very small compared with the large scale of the flow (i.e. ). Since eddies in this
range are much larger than the dissipative eddies that exist at Kolmogorov scales, kinetic
energy is essentially not dissipated in this range, and it is merely transferred to smaller scales
until viscous effects become important as the order of the Kolmogorov scale is approached.
Within this range inertial effects are still much larger than viscous effects, and it is possible to
assume that viscosity does not play a role in their internal dynamics (for this reason this range
is called "inertial range"). Turbulence (modern definition) always cascades from small scales
to large.

Hence, a third hypothesis of Kolmogorov was that at very high Reynolds number the
statistics of scales in the range are universally and uniquely determined by the
scale r and the rate of energy dissipation .

The way in which the kinetic energy is distributed over the multiplicity of scales is a
fundamental characterization of a turbulent flow. For homogeneous turbulence (i.e.,
statistically invariant under translations of the reference frame) this is usually done by means
of the energy spectrum function E(k), where k is the modulus of the wave vector
corresponding to some harmonics in a Fourier representation of the flow velocity field u(x):

Where û(k) is the Fourier transform of the velocity field. Thus, E (k)dk represents the
contribution to the kinetic energy from all the Fourier modes with k < |k| < k + dk, and
therefore,

The wave number k corresponding to length scale r is k = 2π / r. Therefore, by dimensional


analysis, the only possible form for the energy spectrum function according with the third
Kolmogorov's hypothesis is
49
,

Where C would be a universal constant. This is one of the most famous results of
Kolmogorov 1941 theory, and considerable experimental evidence has accumulated that
supports it.

In spite of this success, Kolmogorov theory is at present under revision. This theory
implicitly assumes that the turbulence is statistically self-similar at different scales. This
essentially means that the statistics are scale-invariant in the inertial range. A usual way of
studying turbulent velocity fields is by means of velocity increments:

That is, the difference in velocity between points separated by a vector r (since the turbulence
is assumed isotropic, the velocity increment depends only on the modulus of r). Velocity
increments are useful because they emphasize the effects of scales of the order of the
separation r when statistics are computed. The statistical scale-invariance implies that the
scaling of velocity increments should occur with a unique scaling exponent β, so that when r
is scaled by a factor λ,

Should have the same statistical distribution as

With β independent of the scale r. From this fact, and other results of Kolmogorov 1941
theory, it follows that the statistical moments of the velocity increments (known as structure
functions in turbulence) should scale as

Where the brackets denote the statistical average, and the Cn would be universal constants.

There is considerable evidence that turbulent flows deviate from this behavior. The scaling
exponents deviate from the n/3 value predicted by the theory, becoming a non-linear function
of the order n of the structure function. The universality of the constants have also been
questioned. For low orders the discrepancy with the Kolmogorov n/3 value is very small,
which explain the success of Kolmogorov theory in regards to low order statistical moments.
In particular, it can be shown that when the energy spectrum follows a power law

with 1 < p < 3, the second order structure function has also a power law, with the form

50
Since the experimental values obtained for the second order structure function only deviate
slightly from the 2/3 value predicted by Kolmogorov theory, the value for p is very near to
5/3 (differences are about 2%). Thus the "Kolmogorov -5/3 spectrum" is generally observed
in turbulence. However, for high order structure functions the difference with the
Kolmogorov scaling is significant, and the breakdown of the statistical self-similarity is clear.
This behavior and the lack of universality of the Cn constants are related with the
phenomenon of intermittency in turbulence. This is an important area of research in this field,
and a major goal of the modern theory of turbulence is to understand what is really universal
in the inertial range.

Laminar flow:
Laminar flow, sometimes known as streamline flow, occurs when a fluid flows in parallel
layers, with no disruption between the layers. In fluid dynamics, laminar flow is a flow
regime characterized by high momentum diffusion, low momentum convection, pressure and
velocity independent from time. It is the opposite of turbulent flow. In nonscientific terms
laminar flow is "smooth," while turbulent flow is "rough."

The dimensionless Reynolds number is an important parameter in the equations that describe
whether flow conditions lead to laminar or turbulent flow. Reynolds numbers of less than 500
are generally considered to be of a laminar type. When the Reynolds number is much less
than 1, creeping motion or Stokes flow occurs. This is an extreme case of laminar flow where
viscous (friction) effects are much greater than inertial forces.

For example, consider the flow of air over an airplane wing. The boundary layer is a very thin
sheet of air lying over the surface of the wing (and all other surfaces of the airplane). Because
air has viscosity, this layer of air tends to adhere to the wing. As the wing moves forward
through the air, the boundary layer at first flows smoothly over the streamlined shape of the
airfoil. Here the flow is called laminar and the boundary layer is a laminar layer.

For a practical demonstration of laminar and non-laminar flow, one can observe the smoke
rising off a cigarette in a place where there is no breeze. The smoke from the cigarette will
rise vertically and smoothly for some distance (laminar flow) and then will start undulating
into a turbulent, nonlaminar flow

51

Anda mungkin juga menyukai