Vous êtes sur la page 1sur 9

Equation (4.

47) is the net flux of the extensive property N through the boundary of the control
volume. The product of density, velocity, and area is the mass flow rate. Thus, the equation
really represents the mass flow in and out of the control volume times the associated intensive
property.

The boundary is called the control Surface, or CS. The vector dot product of the velocity and
the area signifies that only the velocity component normal or perpendicular to the surface can
be used for calculation purposes. An example of how this is used practically can be seen from
the conservation of mass as given by:


=0 (4.48)

The associated intensive property for mass is simple one. Inserting the above information into
the general control volume statement yields:

0=

+ . (4.49)

Notice the first term in Eq. (4.49). Practically speaking, this term takes into account the change
in density of the control volume. For steady-state and incompressible analysis, the equation
reduces to:

(Mass in) Control Volume = (Mass out) Control Volume

For constant velocity across any opening, the equation becomes:

()1 = ()2 (4.50)

Equation (4.50) is the standard form for calculating mass balances. Additional examples of the
method are found in Fox and McDonald3 and Potter and Foss.13 Application of the method to
eductors is provided in Chapter 8 of this book. In addition to fixed control volumes, it is
possible to construct control volumes such that all or part of the surfaces move. Required
modifications are outlined in both Panton1 and Foss.14

4.3.4 Differential Formulation

All of the previous fluid flow analysis required various simplifications and assumptions. The
most fundamental formulation of fluid phenomena is the differential formulation. Numerous
excellent references exist (e.g., Panton,1 Potter and Foss,13 and Kuo14) for the reader interested
in greater detail on this subject. This treatment concentrates on the conceptual development
of the formulation and the final form. It is the intent of this reference to provide insight into
the meaning of the equations, how simplifications have been applied to them, and the
limitations of the simplifications.

On a conceptual level, the equations of interest are partial differential equations that describe:

1. Conservation of mass
2. Conservation of momentum
3. Conservation of energy
4. Conservation of species

Similar to density, the main assumption regarding the differential derivation of these
equations is that the volume reduces to a point, but a point much larger than the molecular
length scale. Each of the conservation principles are discussed in turn.
Because it is fairly simple, a reasonably complete derivation of conservation of mass is
provided. Insight gained from the conservation of mass derivation will be generalized as
appropriate for the other conservation equations. Details will be referenced, as appropriate.

4.3.4.1 Conservation of Mass

Figure 4.12 is an idealization of a small differential control volume. For simplicity, Cartesian
coordinates (X, Y, Z) are utilized; however, the results could be generalized to any other
system. The little cube has dimensions of dx by dy by dz, and thus the area of any face is the
same (i.e., dx.dy), and the volume is dx.dy.dz. Reference will be made to these observations in
all of the following sections.

If this small control volume is placed at a fixed location in a fluid flow field, then fluid will flow
into and out of the volume continuously. In other words, conservation of mass states that the
amount of mass going into the volume plus any mass stored in the volume. The mass flow
across any face into the volume is going to be the area of the face, times the velocity, times
the density, or in equation form:

Mass flow through a face = (Velocity component) (Face area) (Density)

From Section 4.2.1, it is observed that the differential volume is chosen so that the density is
uniform throughout, but is not so small that molecular spacing is sparse. Choosing the X-
direction for analysis, as shown in Figure 4.13, results in the following:

(Mass flow)x = U(dy.dz)() (4.51)

where

U = Velocity in the X-direction

= Density

For conservation of mass analysis, however, the point of interest is ensuring that the mass
entering, leaving, and being stored are the same. The real question that the equation must
answer is: What is the difference between mass entering and exiting in the X-direction? or:

(Mass entering)x - (Mass leaving)x = U(dy.dz) - (U + d(U)(dy.dz) = -d(U)(dy.dz) (4.52)

In Eq. (4.52), is simply the change in X-direction velocity that occurred over the distance dx,
so the change in mass flow rate over distance dx is ().

The mass of a cube is the volume times the density. In Eq. (4.52), the rate at which mass in
entering or leaving the cube is represented. This being the case, the conservation of mass for
this system will be a rate equation, that is, how much has the mass changed with time. This is
expressed as:

Mass stored/time = {( Mass at time 1) ( Mass at time 2)} / time =



{ 1(dx . dy. dz) - 2(dx . dy. dz)} / dt = (dx.dy.dz) (4.53)

where

d = Change in density with time

dt = Differential change in time


dx.dy.dz = Volume of cube

The above expression is very reasonable. The only way that mass can be stored in a fixed-
volume container of any sort is for the density to change. Repeating the mass flow analysis for
each of the other two directions, collecting the expressions, and setting them equal to the
expression for the change in mass, results in the following general expression for conservation
of mass for fluid flow:

(. . ) = ()(. ) ()(. ) ()(. ) (4.54)

Dividing Eq. (4.54) by (. . ) and collecting terms on one side of the equation:
() () () ()

+
+
+
=0 (4.55)

where

U = X velocity component

V = Y velocity component

W = Z velocity component

The above expression has the following physical implications:



1. The time-dependent term represents the change of mass inside the volume.
2. The other terms represent the difference between what leaves and enters the volume.
3. This is summarized as:

() () () ()

+
+
+
=0 (4.56)

Internal changes Mass difference between entering and leaving volume

For incompressible flow, this reduces to:


() () () ()
+ + + =0

As it turns out, the general form of Eq. (4.56) can be used for any physical quantity of interest:
momentum, energy, chemical species, etc. In general, this form of equation is said to
represent the time variant and convection terms of the differential equations. In fact, this form
appears so often that it is given the special name of substantial derivative and represented in
shorthand as follows:
( ) ( ) ( ) ( ) ( )

=
+ + + (4.57)

Where different physical parameters can be substituted into the blanck sets of parenthese.
This would allow the conservation of mass for incompressible flow to be written as:
()

=0 (4.58)

Use of the substantial derivative will be routinely referenced in the following sections.
4.3.4.2 Conservation of Momentum

Newtons second law of motion stated in words is: the time rate of change of momentum is
equal to the sum of all applied forces. In equation form, this is:
()

= (4.59)

Instead of change of momentum, many physics textbooks refer to mass times acceleration.
The two statements are equivalent for volumes with constant density; but for fluid
considerations, changes in momentum are more appropriate. Momentum is a vector quantity
with three components, so any equation derived for one direction can be generalized into
three similar equations, one for each component (x,y,z directions). For simplicity, only the X-
directions is condidered here.

The rate of momentum change is represented as: (1) the difference between the momentum
entering the volume and the momentum leaving the volume, and (2) the momentum stored. It
can be shown this is the same type of convection and storage term derived for conservation of
mass, so the substantial derivative can be used to describe a change in momentum. For
momentum, velocity times density is utilized; thus:
() ()

=
(4.60)

Forces that can be applied fall into two categories: (1) forces that are applied to the entire
volume equally, and (2) forces that are applied to the surface of the volume. The first type of
force is sometimes called a body force. Gravity is normally the only body force encountered,
and will be the only one considered herein. Others, however, are possible, such as
electromagnetic fields and acceleration. The second type of force, surface forces, can act
either in a direction normal to the surface (pressure), or in a direction parallel to the surface
(shear stress). The analytical form of each of these forces is summarized below.

Gravitational Body Force: The gravitational body force is illustrated in Figure 4.14 and is
symbolized by Fx. If there is no body force in the X-direction, then this term is ignored.
Common practice is to consider the Y-direction as up and down, so the gravitational body force
term usually would only apply to the Y equation.

Normal or Pressure Forces: Normal or pressure forces are illustrated in Figure 4.15 and take
the following form:

(4.61)

A net force will arise only if there is a difference in the direction of interest. The negative sign
occurs because higher pressures in the direction of interest will induce a flow in the negative
direction.

Parallel or Shear Stress: Parallel or shears stress is illustrated in Figure 4.16. For a Newtonian
fluid, shear stress or force will be proportional to the velocity gradient (see Section 4.4.1.1).
The force due to shear stress along the Y-direction on the X-plane is:

= (4.62)

However, what is of interest is the change in force (the force in this case being shear stress)
along any one direction.
This change in force formulation gives rise to terms that look like:
2
(4.63)
2

Collecting all of the above terms results in the following.

X-direction momentum equation:

In a similar manner, equations for the Y and Z momentum can be derived. The analytical form
of the above equation may at first be quite intimidating. However, it is nothing more than a
logical extension of Newtons second law of motion from high school physics (F=ma), which
has been applied to fluid flows. Strictly speaking, the above equation is only valid for
incompressible flow. The interested reader is directed to previously listed references for
additional details and an extension of the momentum equation to include compressibility
effects.

4.3.4.3 Conservation of Energy

The conservation of energy equation derivation is similar to mass and momentum. First, start
with a general conservation principle, in this case the first law of thermodynamics, as
discussed in Van Wylen:15
(Heat transferred) + (Energy in) + (Energy generated) = (Energy stored) + (Energy out) + (Work) (4.65)

Or, moving terms around:


(Energy stored) + (Energy out) (Energy in) = (Heat transferred) + (Energy generated) (Work) (4.66)

The term on the left will again be the substantial derivative of energy into and out of the
volume. Depending on the processes involved, terms on the right will vary, but may include all
or none of the following:

1. Convective heat transfer (heat transfer)


2. Conductive heat transfer (heat transfer)
3. Radiative heat transfer (heat transfer)
4. Chemical heat release (energy generated)
5. Expansion (work)
6. Wall friction (work)
7. Viscous dissipation (work)

In general, for combustion in the petrochemical industries, the common terms utilized are
convective and radiative heat transfer, chemical heat release, and wall friction. Inclusion of
multiple chemical species complicates the formulation, as molecular diffusion of the various
species provides an energy in a flowing fluid can be separated to consider thermal and kinetic
energy separately. In low Mach-number flows (i.e., incompressible flows), conservation of
kinetic energy is automatically satisfied whenever the momentum equation is satisfied (the
conservation of kinetic energy equation can be derived by manipulating the momentum
equation). The following (thermal) energy equation contains the terms that would normally be
important in a combusting flow (numbers correspond to terms described above):

The reader should consult Bird et al.16 for a much more complete discussion of energy
conservation.

4.3.4.4 Conservation of Species


If there are no chemical reactions, then the implications that the mass fraction of any
particular species can only change due to mass flow into and out of the volume. Under this
constraint, conservation of species becomes exactly like the conservation of mass equation,
namely:

If reactions are occurring, then the change will be equal to the formation (or destruction) rate
of the compound. Reaction rate equations are normally of the Arrhenius type (see Kuo14). This
results in the conservation of species equation being:

4.4 DIFFERENT TYPES OF FLOW

4.4.1 Turbulent and Laminar Flow

For centuries, scientists have been fascinated with flows in nature. Visual observations of fluids
in motion have resulted in two broad classifications of flows: laminar and turbulent. Laminar
flows have no fluctuations and tend to flow for long distances with very little mixing occurring
in the flow. Figures 4.17 and 4.18 illustrate two common examples of laminar flow. In figure
4.17, smoke from burning incense slowly rises in still air for a distance of 6 to 12 in. (15 to 30
cm) before becoming unsteady. In figure 4.18, water exits a water faucet at a very low velocity
and falls in a straight line until it strikes and object or is disrupted by air currents. In both cases,
the regions of smooth movement are examples of laminar flows.

Turbulent flows have fluctuations and eddies associated with them that increase mixing rates
substantially. The legendary artist and scientist Leonardo da Vincis view on turbulence is
reproduced in figure 4.19. As shown, the circulation, induced vortices and fluctuations allow, in
this case, an introduced stream to mix rapidly with a still liquid. Most flows in nature will tend
to become turbulent as flow rate or local velocity is increased.

4.4.1.1 Reynolds Number

During the 1880s, Osborn Reynolds quantified the previous qualitative observations of laminar
and turbulent flows. His observations were first published in 1883.17 A schematic of his
experimental apparatus is shown in Figure 4.20. In this device, water flows from a tank
through a bell-mouthed inlet into a glass pipe. Also, at the entrance to the glass pipe is the
outlet of a small tube, which leads to a reservoir of dye.

Reynolds discovered that at low water velocities, the stream of dye issuing from the thin tube
did not mix with the water. Instead, the dye became a distinct flow, parallel to the pipe
centerline. As the valve was opened and the water velocity increased, it was observed that at
some greater velocity, the dye would rapidly mix with the water, causing the entire flow to be
colored with the dye. Reynolds deduced from these experiments that two flow regimes
existed. In the first (laminar flow), the fluid streams flow past each other in parallel layers, or
laminae, but do not mix with each other. The second regime, at higher velocities, is where the
two streams mix rapidly. This mixing is caused by vortices, illustrated in Figures 4.21 and 4.22,
that arise in the flow. Due to the presence of the vortices, this flow has been called turbulent.
Figure 4.23 illustrates laminar flow.

Reynolds found that below a certain velocity, the flow always became laminar. Figure 4.23
shows laminar flow can even be achieved by flow over a cube. As the velocity was increased,
turbulent flow could always be achieved, but the precise velocity depended on how still the
water in the tank was prior to the experiment. In addition, Reynolds was able to generalize his
results into a no dimensional parameter, which today is known as the Reynolds number, Re. It
is defined as follows:

Where

V = Velocity in pipe

D = Pipe diameter

= Density of fluid

= Fluid viscosity

As a practical matter, pipe flows having a Reynolds number less than 2300 are laminar, and
flows having a Reynolds number greater than 4000 are turbulent. In addition to determining
type of flow, Reynolds numbers have been proven to also scale the intensity of turbulence in a
flow. That is, higher Reynolds numbers result in greater vortex generation and faster mixing
rates. As a result, Reynolds number calculations are very common in the petrochemical
industry. They are utilized to scale orifice coefficients,19 friction factors (Vernard and Street,4
and Section 4.5 of this book), heat transfer rates (Lienhard,19 and Chapter 3 of this book), and
mass transfer rates (Bird16).

4.4.1.2 Hydraulic Diameter

In addition to circular pipe flows, Reynolds numbers can be utilized, as described above, for
geometries other than round. Square, triangular, and other enclosed flows scale with Reynolds
numbers. For geometric variations with no obvious diameter, a concept termed the
hydraulic diameter is useful. Essentially, it is the ratio of the flow area to the flow perimeter,
multiplied by four. In analytical form, this ratio is expressed as:

It should be noted that the above definition of hydraulic diameter reduces to the definition of
normal diameter for circular pipes. The above equation provides reasonable accuracy for
calculations involving turbulent flow, but large errors will occur for laminar flow calculations.
See Vernard and Street4 for additional details.

4.4.1.3 Reynolds Averaging

Turbulent flows are characterized by a highly fluctuating instantaneous velocity. The flow fields
generated are three-dimensional in nature and dependent on an unknown time function. To
help understand the processes, Reynolds34 suggested utilizing a velocity composed of two
components: (1) a time averaged velocity, and (2) a fluctuating velocity. This is a logical
substitution, because, for 99% of applications, only the mean (or average) velocity is of
interest, and the time average pf the fluctuating velocity would be zero. This is expressed as:

The two velocity terms can be substituted into the differential momentum equation (4.64),
and the resulting equation averaged over time to obtain an expression for average velocity. It
should be noted that while the time average of fluctuating velocity is zero, the time average of
the product of two fluctuating velocities is not zero. Potter and Foss13 present a good
discussion of this observation.

Starting from Eq. (4.64) (the momentum equation), if body forces are neglected and density is
considered constant, the x only equation can be rewritten as:

Where
The part of the momentum equation contains no cross velocity terms. However, close
examination of the substantial derivative (see Section 4.3.4 of this chapter) reveals that the
left-hand side of Eq. (4.73) does contain the products of different velocities. Substitution of the
decomposed velocity into the momentum equation, taking the time average (to develop a
time-averaged momentum equation), and simplifying (see Potter and Foss13 for details),
results in the following x momentum equation:

Interestingly, the cross fluctuating velocity terms have the same derivative operator as the
shear stress, so it is common to simply redefine the shear stress with those terms added to it.
In that case, the momentum equation for time-averaged velocity will have the exact same
form as before, namely:

But the stress term for the x equation is now modified by the fluctuating velocity terms. It now
has the following form:

It should be noted that although the fluctuating terms are historically lumped with the stress
and pressure terms (forces, from the derivation), the actually arise from the convective side of
the momentum equation. If the same procedure is followed for the y and z momentum
equations, a total of nine fluctuating velocities arise from the derivation. Noting, however, that
velocity order can be interchanged, there are only six different terms. Collecting them in
tensor form (terms from the three different equations are listed together), they are as follows:

Collectively, these terms are called Reynolds stresses. The term stress is applied to them
because they are associated with the stress term in the momentum equation, although they
arise from the convective side of the momentum equation. Research over the last 100 years
has not yet provided a completely satisfactory model for these terms. Computational fluids
dynamics (CFD, see Chapter 9 of this book) makes extensive use of various assumption and
modeling approximations. Occasionally, the ability to accurately model a flow field still
depends on the modelers understanding of the limitations of the chosen Reynolds stress
model.

4.4.1.4 Jets

When a fluid emerges from a nozzle, it will interact with the surrounding fluids. This type of
system, termed a free jet, commonly occurs in combustion systems (Figure 4.24). High-
pressure fuel from a nozzle, steam spargers, and liquid fuel sprays are all examples of free jets.
Figure 4.25 illustrates the interaction of a free jet with the surrounding fluid. As the jet travels
downstream from the nozzle, its diameter will increase as it captures ambient fluid into its
stream. This phenomenon has been extensively studied, both experimentally and theoretically.
For theoretical treatment, the interested reader is referred to Schlichting,20 Hinze,21and
Tennekes.22 The historical treatment is to assume that fluctuating velocities have the same
magnitude in each of the three directions (this is called isotropic turbulence), which reduces
the Reynolds stresses from six terms to one term. The single stress term is then modeled as a
function of the local velocity gradient and the nozzle diameter. This model has a single fitted
parameter, which is deduced from experimental data.

Experimental characterization of nozzles has resulted in practically the same results as those
obtained theoretically, except near-nozzle effects are quantified. Nevertheless, both
treatments result in the following two equations describing jet velocities (from Beer and
Chigier23):

Vous aimerez peut-être aussi