You are on page 1of 70

Types of Fluid Flow

Types of Fluid Flow


NEWTON'S LAW OF VISCOSITY (MOLECULAR TRANSPORT OF MOMENTUM)

τyx force in the x direction on a unit


area perpendicular to the у direction
or the force exerted by the fluid of
lesser у on the fluid of greater y.

1.1

➢ The flow is called 'laminar" because the adjacent layers of fluid ("laminae") slide past one another in an
orderly fashion when syrup is poured, in contrast to "turbulent" flow, which is the irregular, chaotic flow one
sees in a high-speed mixer.
NEWTON'S LAW OF VISCOSITY contd -------------
𝑑𝑉𝑥
τyx = - μ 1.2
𝑑𝑦

This equation, which states that the shearing force per unit area is proportional to the negative of the velocity gradient, is often called
Newton's law of viscosity. In another fashion, τyx may also be interpreted as the flux of x-momentum in the positive у direction under
steady state.
➢ Newtonian fluids: - resistance to flow of all gases and all liquids with molecular weight of less than about
5000 whereas Polymeric liquids, suspensions, pastes, slurries, and other complex fluids are not described by
Eq. 1.2 and are referred to as non-Newtonian fluids.
μ
kinematic viscosity ν = ρ

Table 1.1 Unit of quantities related to EQ. 1.2 Table 1.2 Viscosity of Water and Air at 1 atm Pressure
NEWTON'S LAW OF VISCOSITY (MOLECULAR TRANSPORT OF MOMENTUM)

Table 1.4 Viscosity of some liquid metals

Table 1.3 Viscosity of Some Gases and Liquids at Atmospheric Pressure

➢ Note that for gases at low density, the viscosity increases with increasing temperature, whereas for liquids
the viscosity usually decreases with increasing temperature.
Ex. 1.1 Compute the steady-state momentum flux τух in lbf/ft2 when the lower plate velocity V in above Fig. 1-1 is 1 ft/s in the positive x
direction, the plate separation У is 0.001 ft, and the fluid viscosity µ is 0.7 cp.

GENERALIZATION OF NEWTON'S LAW OF VISCOSITY


To do this we consider a very general flow pattern, in which the fluid velocity may be in various directions at various places and may depend on
the time t. The velocity components are then given by V = V (vx, vy, vz, t) where,
vx = vx(x, y, z, t); vy = vy(x, y, z, t); vz = vz(x, y, z, t) 1.3

In such a situation, there will be nine stress components τij(where i and j may take on the designations x, y, and z), instead of the component τух
that appears in Eq. 1.2. We therefore must begin by defining these stress components.

Two contributions to the force: -


The pressure, and The viscous forces

Fig. 1.2 Pressure and viscous forces acting on planes in the fluid perpendicular to the three coordinate systems. small cube-shaped volume element within the flow field, each
face having unit area. The center of the volume element is at the position x, y, z.
GENERALIZATION OF NEWTON'S LAW OF VISCOSITY
1. The pressure force (thermodynamic force) will always be perpendicular to the exposed surface. Hence in (a) the force per unit area on the shaded surface

will be a vector pδx—that is, the pressure (a scalar) multiplied by the unit vector δx in the x direction. Similarly, the force on the shaded surface in (b) will

be pδy , and in (c) the force will be pδz . The pressure forces will be exerted when the fluid is stationary as well as when it is in motion.

➢ The viscous forces come into play only when there are velocity gradients within the fluid.
In general they are neither perpendicular to the surface element nor parallel to it, but rather at some angle to the surface (see Fig. 1.2). In (a) we see a

force per unit area τx exerted on the shaded area, and in (b) and (c) we see forces per unit area τy and τz . Each of these forces (which are vectors) has

components; for example, τx has components τxx, τxy , and τxz.

Therefore, for Fig 1.2 (a) Viscous Force τx can be written as summation of τxx along X – direction , τxy along Y – direction andτxz along Z – direction.

Similarly for Fig 1.2 (b) Viscous Force τy can be written as summation of τyx along X – direction , τyy along Y – direction and τyz along Z – direction &

for Fig 1.2 (c) Viscous Force τz can be written as summation of τzx along X – direction , τzy along Y – direction and τzz along Z – direction.

Sometimes we will find it convenient to have a symbol that includes both thermodynamic pressure and the viscous stresses, and so we define the Molecular
stresses as follows
Molecular Stress πij = pδij + τij where i and j may be x, y, or z 1.4

Here δij is the Kronecker delta, which is 1 if i = j and zero if i ≠ j.


GENERALIZATION OF NEWTON'S LAW OF VISCOSITY
Just as in the previous section, the τij (and also the πi) may be interpreted in two way
πij = pδij + τij = force in the j direction on a unit area perpendicular to the i direction or
= flux of j momentum in positive i direction—that is, from the region of lesser x to that of greater x
The stresses πxx = p + τxx , πyy = p + τyy and πzz = p + τzz (Total three) are known as Normal stresses, whereas
the remaining quantities, πxy = τxy, πxz = τxz , πyx = τyx , πyz = τyz , πzx = τzx , and πzy = τzy (Total Six) are called shear stresses.

Table 1.5 Summary of the Components of the Molecular Stress Tensor (or Molecular Momentum-Flux Tensor)"

NOTE: - These quantities, which have two subscripts associated with the coordinate directions, are referred to as "tensors," just as quantities (such as velocity) that have one
subscript associated with the coordinate directions are called “vectors”.
NOTE: - Molecular Momentum-Flux Tensor " because they are associated with the molecular motions. The additional "convective momentum flux tensor" components,
associated with bulk movement of the fluid, will be discussed in later stage of the course.
GENERALIZATION OF NEWTON'S LAW OF VISCOSITY
Navier, Poisson, and Stokes generalized the Newton's law of viscosity equation (1.2) by giving a relation

(1.5)

This shows that there are two coefficients characterizing the fluid: the viscosity µ and the dilatational viscosity κ.
If the fluid is a gas, we often assume it to act as an ideal monoatomic gas, for which к is identically zero. If the fluid is a liquid, we often assume
that it is incompressible, and for incompressible liquids the term containing к is discarded anyway. Knowledge of the volume viscosity is important
for understanding a variety of fluid phenomena, including sound attenuation in polyatomic gases, propagation of shock waves, and dynamics of
liquids containing gas bubbles.

Therefore, we may write

𝜕𝑣𝑗 𝜕𝑣𝑖
τij = - μ ( + ) (1.6)
𝜕𝑥𝑖 𝜕𝑥𝑗

❑ Note on the Sign Convention


GENERALIZATION OF NEWTON'S LAW OF VISCOSITY

𝜕𝑣𝑗 𝜕𝑣𝑖
τij = - μ ( + )
𝜕𝑥𝑖 𝜕𝑥𝑗

Or, Nine stress component can be written in the matrix form as

𝜕𝑣𝑥 𝜕𝑣𝑦 𝜕𝑣𝑥 𝜕𝑣𝑧 𝜕𝑣𝑥


− 2μ −μ ( + ) −μ ( + )
τxx τyx τzx 𝜕𝑥 𝜕𝑥 𝜕𝑦 𝜕𝑥 𝜕𝑧
𝜕𝑣𝑦 𝜕𝑣𝑥 𝜕𝑣𝑦 𝜕𝑣𝑧 𝜕𝑣𝑦
τx𝑦 τyy τz𝑦 = −μ( + ) − 2μ −μ ( + ) (1.7)
𝜕𝑥 𝜕𝑦 𝜕𝑦 𝜕𝑦 𝜕𝑧
τx𝑧 τy𝑧 τzz 𝜕𝑣𝑧 𝜕𝑣𝑥 𝜕𝑣𝑧 𝜕𝑣𝑦 𝜕𝑣𝑧
−μ( + ) −μ ( + ) − 2μ
𝜕𝑥 𝜕𝑧 𝜕𝑦 𝜕𝑧 𝜕𝑧
PRESSURE AND TEMPERATURE DEPENDENCE OF VISCOSITY

The reduced viscosity µr = µൗµc is plotted versus the reduced temperature Tr = 𝑇ൗT
c
𝑝
for various values of the reduced pressure pr = ൗpc. A "reduced" quantity is one
that has been made dimensionless by dividing by the corresponding quantity at the
critical point. The chart shows that the viscosity of a gas approaches a limit (the
low-density limit) as the pressure becomes smaller; for most gases, this limit is
nearly attained at 1 atm pressure.

➢ The viscosity of a gas at low density increases with

increasing temperature, whereas the viscosity of a liquid

decreases with increasing temperature

Fig. 1.3 Reduced viscosity µr = µൗµc as a function of reduced


temperature for several values of the reduced pressure of a gas.
PRESSURE AND TEMPERATURE DEPENDENCE OF VISCOSITY
Ex. 1.2

Where, M molecular weight,


pc critical pressure in atm, Tc critical
temperature in kelvin and µ c in
micropoise

From Fig. 1.3 we obtain µr = µൗµc= 2.39. Hence, the predicted value of the viscosity
µ = µc (µൗµc ) = (189 X 10-6)(2.39) = 452 X 10-6 poise
Viscosity of slag

The liquid oxide slag consists of cations and anions formed by acid – base reactions among the constituents.

An oxide is considered to be acidic, if, when dissolved in the molten slag, it acquires additional oxygen ions to form complex anions

whereas a basic oxide ionizes to contribute oxygen ions to the melt.

The most common acidic constituents in oxide slags is silica (SiO2), and the viscosity of the liquid slag may be discussed by taking

into account the viscosity of pure liquid SiO2 and its variation with the addition of basic oxides like lime (CaO), magnesia (MgO),

sodium oxide (Na2O) etc.

The viscosity of pure liquid SiO2 is very high, of the order of 105 poise. Other oxides like FeO, TiO2, P2O5, Al2O3, Na2O etc.

The viscosity of a slag decreases as the temperature increases. This is due to the fact that, in liquid slags, the ions become more

mobile when the temperature increases. Therefore, the resistance to flow decreases causing a decrease in viscosity.
CONVECTIVE MOMENTUM TRANSPORT
Thus far we have discussed the molecular transport of momentum, and this led to a set of quantities πij, which give the flux of j momentum across a
surface perpendicular to the i direction. We then related the πij to the velocity gradients and the pressure, and we found that this relation involved two
material parameters µ and к
Momentum can, in addition, be transported by the bulk flow of the fluid, and this process is called convective transport. To discuss this we use Fig. 1.4
and focus our attention on a cube-shaped region in space through which the fluid is flowing. At the center of the cube (located at x, y, z) the fluid
velocity vector is V.

Fig. 1.4 The convective momentum fluxes through planes of unit area perpendicular to the coordinate directions
NEWTON'S LAW OF VISCOSITY (MOLECULAR TRANSPORT OF MOMENTUM)

The volume rate of flow across the shaded unit area in (a) is vx . This fluid carries with it momentum ρV per unit
volume. Hence the momentum flux across the shaded area is vx ρV; note that this is the momentum flux from the
region of lesser x to the region of greater x.
Similarly the momentum flux across the shaded area in (b) is vy ρV, and
the momentum flux across the shaded area in (c) is vz ρV.
These three vectors— vx ρV, vy ρV, and vz ρV—describe the momentum flux across the three areas perpendicular to
the respective axes. Each of these vectors has an x, y, and z-component.

Table 1.6 Summary of the Convective Momentum Flux Components


Combined Momentum flux
It is the sum of the molecular momentum flux and the convective momentum flux:
ф = π + ρvv = pδ + τ + ρvv

Keep in mind that the contribution pδ contains no velocity, only the pressure; the combination ρvv

contains the density and products of the velocity components; and the contribution τ contains the

viscosity and, for a Newtonian fluid, is linear in the velocity gradients.

All these quantities are second-order tensors.

фxy = the combined flux of y- momentum across a surface perpendicular to the x direction by molecular

and convective mechanisms.

The second index gives the component of momentum being transported and the first index gives the

direction of transport.
Momentum Flux

Table 1.7 Summary of Notation for Momentum Fluxes

Symbol Meaning

ρvv Convective momentum-flux tensor

τ Viscous momentum-flux tensor

π = pδ + τ Molecular momentum-flux tensor

ф = π + ρvv = pδ + τ + ρvv Combined momentum-flux tensor


Problems

1. Compare Newton's law of viscosity and Hooke's law of elasticity. What is the origin of these "laws"?

2. Verify that "momentum per unit area per unit time'7 has the same dimensions as "force per unit area."

3. Compare and contrast the molecular and convective mechanisms for momentum transport.

4. How do the viscosities of liquids and low-density gases depend on the temperature and pressure?
Chapter 2 Fluid Mechanics
fluid mechanics is defined as the science that deals with the behavior of fluids at rest (fluid statics) or in motion (fluid
dynamics), and the interaction of fluids with solids or other fluids at the boundaries.
What Is a Fluid?
➢ A substance exists in three primary phases: solid, liquid, and gas. (At very high temperatures, it also exists as plasma.).
➢ A substance in the liquid or gas phase is referred to as a fluid.
➢ Distinction between a solid and a fluid is made on the basis of the substance’s ability to resist an applied shear (or
tangential) stress that tends to change its shape.
➢ A solid can resist an applied shear stress by deforming, whereas a fluid deforms continuously under the influence of a
shear stress, no matter how small. In solids, stress is proportional to strain, but in fluids, stress is proportional to strain
rate.
➢ When a constant shear force is applied, a solid eventually stops deforming at some fixed strain angle, whereas a fluid
never stops deforming and approaches a constant rate of strain.

Fig 2.1 Deformation of a rubber block placed between two


parallel plates under the influence of a shear force.
NEWTON'S LAW OF VISCOSITY

Fig 2.2 The behavior of a fluid in laminar flow between two parallel plates when
the upper plate moves with a constant velocity.

In steady laminar flow, the fluid velocity between the plates varies linearly between 0 and V, and thus the
velocity profile and the velocity gradient are

(2.1)

where y is the vertical distance from the lower plate.


NEWTON'S LAW OF VISCOSITY
The angular displacement or deformation (or shear strain) can be expressed as

(2.2)
Rearranging, the rate of deformation under the influence of shear stress τ becomes

(2.3)

Thus we conclude that the rate of deformation of a fluid element is equivalent to the velocity gradient du/dy.
Further, it can be verified experimentally that for most fluids the rate of deformation (and thus the velocity
gradient) is directly proportional to the shear stress τ,

(2.4)

Fluids for which the rate of deformation is linearly proportional to the shear stress are called Newtonian
fluids.
Newtonian & Non-Newtonian Fluid

Fig 2.3 The rate of deformation of a Newtonian fluid


is proportional to shear stress, and the constant of Fig 2.4 Variation of shear stress with the rate of deformation for
proportionality is the viscosity. Newtonian and non-Newtonian fluids.
Note that viscosity is independent of the rate of
deformation for Newtonian fluids.
NON NEWTONIAN FLUIDS
A Non-Newtonian fluid is a fluid that does not follow Newton's law of viscosity, i.e. constant viscosity independent
of stress.
The equation for non Newtonian Fluids is
τ = A*(du/dy)^n + B
Here 'A' represents flow consistency index
'n' represents flow behaviour index
'B' represents yield shear stress
If in this equation when n=1 & B=0 then it became Newtonian fluid.

Let's study the different cases of non newtonian fluid


Case 1 : VARIATION OF VISCOSITY (μ) WITH RESPECT TO RATE OF SHEAR STRESS (dβ/dt)
❑ Dilatant fluid - In dilatant fluids viscosity will increase with respect to rate of shear stress, therefore the slope
of graph will be of increasing nature. Dilatant fluid are shear thickening fluid for example water and sugar
solution. For these fluids value of flow behaviour Index (n) is positive and yield shear stress (B) is zero. Graph
of dilatant fluid are shown in figure below

Fig 2.5 Dilatant liquid


NON NEWTONIAN FLUIDS
❑ Pseudo plastic fluids - In pseudo plastic fluids viscosity will decrease with respect to rate of shear stress,
therefore the slope of graph will be of decreasing nature. Pseudoplastic fluid are shear thinning fluid for example
paint, blood, lipstick. For these Fluids value of flow behaviour index (n) is negative and yield shear stress (B) is
zero.
❑ Bingham fluids - slope of craft will remain constant but the line of graph will not intersect the origin point. Here
flow behaviour index (n) is equal to 1 and yield shear stress (B) will not be equal to zero. Some materials such as
toothpaste and gel can resist a finite shear stress and thus behave as a solid, but deform continuously when the
shear stress exceeds the yield stress and behave as a fluid.

Fig 2.6 Pseudo plastic liquid Fig 2.7 Bingham plastic


NON NEWTONIAN FLUIDS
Case 2 : VARIATION OF VISCOSITY (μ) WITH RESPECT TO TIME
❑ Rheopectic fluids - In Rheopectic fluids viscosity will increase with respect to time. For these Fluids value of
flow behaviour index (n) is positive and yield shear stress (B) is not equal to zero.
❑ Thixotropic fluids - In thixotropic fluids viscosity will decrease with respect to time. For these fluids value of
flow behaviour index (n) is negative and yield shear stress (B) is not equal to zero.

Fig 2.8 Fig 2.9


Fig 2.10 The variation of dynamic (absolute) viscosity of common fluids with temperature at 1 atm
Chapter 3 Shell Momentum Balances and Velocity Distributions in Laminar Flow

Fig. 3-1 (a) Laminar flow, in which fluid layers move


smoothly over one another in the direction of flow, and
(b) turbulent flow, in which the flow pattern is complex and
time-dependent, with considerable motion perpendicular to
the principal flow direction.

SHELL MOMENTUM BALANCES AND BOUNDARY CONDITIONS


For steady flow, the momentum balance is
Total Momentum + ∑ 𝐹 + 0

Conductive Transport of Momentum + Convective Transport of Momentum

rate of momentum in - rate of momentum 𝑜𝑢𝑡 + rate of momentum in - rate of momentum 𝑜𝑢𝑡 +∑𝐹 +0 (3.1)

NOTE: - In shell momentum balance, we apply this statement only to one component of the momentum—namely, the component in the direction of flow.
The procedure in this chapter for setting up and solving viscous flow problems

• Identify the non-vanishing velocity component and the spatial variable on which it depends.

• Write a momentum balance of the form of Eq. 3.1 over a thin shell perpendicular to the relevant spatial variable.

• Let the thickness of the shell approach zero and make use of the definition of the first derivative to obtain the

corresponding differential equation for the momentum flux.

• Integrate this equation to get the momentum-flux distribution.

• Insert Newton's law of viscosity and obtain a differential equation for the velocity.

• Integrate this equation to get the velocity distribution.

• Use the velocity distribution to get other quantities, such as the maximum velocity, average velocity, or force on solid

surfaces.
Boundary Condition at the interface

The most commonly used boundary conditions are as follows:


Gas a. At solid-fluid interfaces the fluid velocity equals the velocity with
τ𝑥𝑦 which the solid surface is moving "no-slip condition."
Plate 1
y b. At a liquid-gas interfacial plane of constant x, the stress-tensor
δ Liquid
z components τxy and τxz are taken to be zero, provided that the gas-side
x L Plate 2
velocity gradient is not too large. This is reasonable, since the viscosities
Fig. 3.1 of gases are much less than those of liquids.

NOTE: - In all of these boundary conditions it is presumed that there is no material passing through the interface; that is, there is no adsorption,
absorption, dissolution, evaporation, melting, or chemical reaction at the surface between the two phases.
FLOW OF A FALLING FILM
Flow of a liquid down an inclined flat plate of length L and width W

➢ Neglecting the disturbances at the edges of the system (z = 0, z = L, у = 0, у = W) if W and L are large compared to t he film thickness δ.

➢ For small flow rates we expect that the viscous forces will prevent continued acceleration of the liquid down the wall, so that vz will
become independent of z in a short distance down the plate.

➢ Therefore, Vz = Vz(x), Vx = 0, and Vy = 0

𝑑𝑉𝑧
➢ Only nonvanishing components of τ are then τXZ = τZX = - μ ( )
𝑑𝑥

Fig. 3.2
Contd-----
Select as the "system" a thin shell perpendicular to the x direction (see Fig. 2.3). Then we set up a Z-momentum balance over this shell, which is
a region of thickness Δx, bounded by the planes z = 0 and z = L, and extending a distance W in the у direction.

z
x

Fig. 3.3

Shell of thickness Δ x over which a z-momentum balance is made.


Arrows show the momentum fluxes associated with the surfaces of the shell.
Since vx and vy are both zero, ρvx vz and ρvy vz are zero. Since vz does not depend on у and z, it follows that τYZ = 0 and τZZ = 0. Therefore, the
dashed-underlined (-----) fluxes in fig 3.3 do not need to be considered. Both ρvz vz are the same at z = 0 and z = L, and therefore do not appear
in the final equation for the balance of z-momentum, Eq. 2.2.
Contd-----
Rate of z-momentum in across surface at z = 0 (WΔx) ϕzz|z=0
rate of z-momentum out across surface at z = L (WΔx) ϕzz|z=L

rate of z-momentum in across surface at x (LW) ϕxz|x


rate of 2-momentum out across surface at x + Ax (LW) ϕxz|x+Δx

gravity force acting on fluid in the z direction (L W Δx)(ρg cos β)


When these terms are substituted into the z-momentum balance of Eq. 3.1
{(LW) ϕxz|x - (LW) ϕxz|x+Δx } + {(WΔx) ϕzz|z=0 - (WΔx) ϕzz|z=L} + (L W Δx)(ρg cos β) = 0
(3.2)
Divide both the sides by LWΔx, and the limit taken as Δx approaches zero
ϕ | −ϕxz|x ϕzz|z=0 −ϕzz|z=L
lim { xz x+Δx } – { } = ρg cos β (3.3)
Δx→0 Δx Δx

𝜕ϕxz ϕzz|z=0 −ϕzz|z=L


➢ -{ } = ρg cos β (3.4)
𝜕𝑥 Δx
𝜕𝑣
We know, ϕxz = τxz + ρvxvz = - μ 𝜕𝑥𝑧 + ρvxvz (3.5)
𝜕𝑣𝑧
ϕzz = τzz + ρvzvz = 2μ + ρvzvz (3.6)
𝜕𝑧
Contd-----
➢ A/c to the postulates Vz = Vz(x), Vx = 0, and Vy = 0, we see that
(i) since Vx = 0, the ρvxvz term in Eq. 3.5 is zero;
𝜕𝑣𝑧
(ii) since Vz = Vz(x), the term 2μ in Eq. 3.6 is zero; and
𝜕𝑧

(iii) since Vz = Vz(x), the term ρvzvz is the same at z = 0 and z = L.

𝑑τ
Hence τxz depends only on x, xz = ρg cos β (3.7)
𝑑𝑥
This is differential equation for Momentum flux.
➢ τxz = ρg cos β x + C1 Now, Use the Boundary condition : - at liq-gas interface, x = 0, τxz = 0
➢ Therefore the momentum-flux distribution is τxz = (ρg cos β) x

𝜕𝑣𝑧
➢ From Newton’s law of viscosity, τxz = - μ = (ρg cos β) x
𝜕𝑥 ρgδ2cos β 𝑥 2
𝜕𝑣𝑧 ρg cos β Vz = [1- (δ) ] (3.8)
➢ = - ( ) x 2μ
𝜕𝑥 μ
ρg cos β 2
➢ Vz = - ( ) x + C2

➢ Boundary condition : - no-slip condition at solid - liquid interface, at x = δ, Vz = 0
ρg cos β
➢ C2 = ( 2μ ) δ2
Average velocity, Maximum velocity & the force on the plate exerted by liquid
ρgδ2cos β
❖ The maximum velocity Vz, max is the velocity at x = 0 is (3.9)

❖ The average velocity <vz> over a cross section of the film is obtained as follows

𝑤 δ
‫׬‬0 ‫׬‬0 𝑉𝑧 𝑑𝑥 𝑑𝑦 1 δ
<vz> = 𝑤 δ = ‫׬‬0 𝑉𝑧 𝑑𝑥
δ
volume flow rate/cross-sectional area
‫׬‬0 ‫׬‬0 𝑑𝑥 𝑑𝑦

ρgδ2cos β 2
(3.10)
<vz> = = Vz, max
3μ 3

𝑤 δ ρ2gwδ3 cos β
❖ The mass rate of flow ω = ρ × volume flow rate = ‫׬‬0 ‫׬‬0 ρ 𝑉𝑧 𝑑𝑥 𝑑𝑦 = ρ wδ <vz> = (3.11)

❖ The z-component of the force F of the fluid on the solid surface is obtained by integrating the shear stress over the
fluid-solid interface:
𝐿 𝑊 𝐿 𝑊 𝜕𝑣𝑧 ρgδ cos β
➢ Fz = ‫׬‬0 ‫׬‬0 (τxz|x=δ )𝑑𝑦 𝑑𝑧 = ‫׬‬0 ‫׬‬0 (−μ |x=δ )𝑑𝑦 𝑑𝑧 = (LW) (−μ) - ( )
𝜕𝑥 μ

➢ Fz = ρgδ LW cos β (3.12)

❑ This is the z-component of the weight of the fluid in the entire film
NEWTON'S LAW OF VISCOSITY (MOLECULAR TRANSPORT OF MOMENTUM)

4 δ <vz> ρ
For falling films the Reynolds number is defined by Re = .
μ

The three flow regime are then:

➢ laminar flow with negligible rippling Re < 20

➢ laminar flow with pronounced rippling 20 < Re < 1500

➢ turbulent flow Re > 1500

NOTE : - The above analysis we have derives above is valid only for the first regime.
Momentum-flux distribution and the velocity distribution in a falling film
Chapter 4 The Equations of Change for Isothermal Systems
❑ The equation of continuity (for the mass balance)

❑ The equation of motion (for the momentum balance)

❑ The equation of mechanical energy

❑ The equations of change in terms of the substantial derivative

➢ NOTE : - These equations can be used as the starting point for studying all problems involving the isothermal flow of a pure fluid.
➢ The equation of continuity is developed by making a mass balance over a small element of volume through which the fluid is
flowing. Then the size of this element is allowed to go to zero (thereby treating the fluid as a continuum), and the desired partial
differential equation is generated.
➢ The equation of motion is developed by making a momentum balance over a small element of volume and letting the volume
element become infinitesimally small. Here again a partial differential equation is generated. This equation of motion can be used,
along with some help from the equation of continuity, to set up and solve all the problems given in Chapter 3.
➢ "substantial derivative." This is the time derivative following the motion of the substance (i.e., the fluid).
THE EQUATION OF CONTINUITY
This equation is developed by writing a mass balance over a volume element Δx Δy Δz, fixed in space, through
which a fluid is flowing

{rate of increase of mass} = {rate of mass in } - {rate of mass out} (4.1)

We begin by considering the two shaded faces, which are perpendicular to the x-axis.

Fig. 4.1
THE EQUATION OF CONTINUITY

(4.2)

(4.3)

(4.4)
Now total rate of mass flow to the control volume

(4.5)

(4.6)

This equation is called the continuity equation in Cartesian coordinates.

In vector form (4.7)

In cylindrical polar coordinates the continuity equation can be written as

(4.8)
(4.7)

For incompressible flow

(4.9)
NEWTON'S LAW OF VISCOSITY (MOLECULAR TRANSPORT OF MOMENTUM)

(4.10)

(4.11)

Here is called the "divergence of ρv," sometimes written as "div ρv." The vector ρv is the mass flux, and its
divergence has a simple meaning: it is the net rate of mass efflux per unit volume.

A very important special form of the equation of continuity is that for a fluid of constant density
(incompressible fluid) ( • v) = 0

Ex. Show that for any kind of flow pattern, the normal stresses are zero at fluid-solid boundaries, for Newtonian fluids
with constant density. This is an important result that we shall use often.
THE EQUATION OF MOTION
This equation is developed by writing a momentum balance over a volume element Δx Δy Δz, fixed in space, through which a
fluid is flowing
{rate of increase of momentum} = {rate of momentum in } - {rate of momentum out} + {External force on fluid} (4.12)

Fig. 4.2. Fixed volume element Δx Δy Δz , with


six arrows indicating the flux of x-momentum
through the surfaces by all mechanisms. The
shaded faces are located at x and x + Δx.

Note that Eq. 4.12 is an extension of Eq. 3.1 to unsteady-state problems. However, in addition to including the unsteady-state term, we must allow the fluid
to move through all six faces of the volume element. Remember that Eq. 4.12 is a vector equation with components in each of the three coordinate
directions x, y, and z. We develop the x-component of each term in Eq. 4.12; the y- and z-components may be treated analogously.
THE EQUATION OF MOTION
First, we consider the rates of flow of the x-component of momentum into and out of the volume element shown in
Fig. 4.2. Momentum enters and leaves Δx Δy Δz by two mechanisms: convective transport and molecular transport.

enters across the face at x = (фхх) х Δy Δz


x- component of momentum
Leaves the face at x+Δx = (фхх) х+Δx Δy Δz

enters across the face at y = (фyх) y Δz Δx


x- component of momentum
Leaves the face at y+Δy = (фyх) y+Δy Δz Δx

enters across the face at z = (фzх) z Δx Δy


x- component of momentum
Leaves the face at z+Δz = (фzх) z+Δz Δx Δy

When these contributions are added we get for the net rate of addition of x-momentum

Δy Δz {(фхх) х - (фхх) х+Δx} + Δz Δx {(фyх) y - (фyх) y+Δy} + Δx Δy {(фzх) z - (фzх) z+Δz} (4.13)

across all three pairs of faces.


THE EQUATION OF MOTION
Next there is the external force (typically the gravitational force) acting on the fluid in the volume element.

The x-component of this force is ρgx Δx Δy Δz (4.14) Assume 𝑔 = gx 𝑖 + gy 𝑗 + gz 𝑘

Equations 4.13 and 4.14 give the x-components of the three terms on the right side of Eq. 4.12.

The sum of these terms must then be equated to the rate of increase of x-momentum within the volume element:
𝜕ρ𝑣𝑥
Δx Δy Δz . (4.15)
𝜕𝑡

Now put all the terms in eq. 4.12 for x- direction momentum

{rate of increase of momentum} = {rate of momentum in } - {rate of momentum out} + {External force on fluid}

𝜕ρ𝑣𝑥
Δx Δy Δz = Δy Δz {(фхх) х - (фхх) х+Δx} + Δz Δx {(фyх) y - (фyх) y+Δy} + Δx Δy {(фzх) z - (фzх) z+Δz} + ρgx Δx Δy Δz
𝜕𝑡
Divide both sides by Δx Δy Δz and the limit is taken as Δx, Δy, and Δz go to zero
𝜕ρ𝑣𝑥 𝜕ф 𝜕ф 𝜕ф
= - ( хх + yх + zх) + ρgx
𝜕𝑡 𝜕𝑥 𝜕𝑦 𝜕𝑧
Similar equations can be developed for the y- and z-components of the momentum balance:
𝜕ρ𝑣𝑦 𝜕ф 𝜕ф 𝜕ф
= - ( хy + yy + zy) + ρgy and
𝜕𝑡 𝜕𝑥 𝜕𝑦 𝜕𝑧

𝜕ρ𝑣𝑧 𝜕 фх 𝑧 𝜕фyz 𝜕фzz


=-( + + ) + ρgz
𝜕𝑡 𝜕𝑥 𝜕𝑦 𝜕𝑧
𝜕ρ𝑣𝑖
By using vector-tensor notation, these three equations can be written as follows: = - [ .φ]i + ρgi where i can be x, y, and z. (4.16)
𝜕𝑡
THE EQUATION OF MOTION
The quantities ρ𝑣 i { are the Cartesian components of the vector ρ𝑣, which is the momentum per unit volume at a point in the fluid.
Similarly, the quantities ρgi are the components of the vector ρg, which is the external force per unit volume. The term —[ • ф]i, is the
ith component of the vector -[ • ф].
When the ith component of Eq. 4.16 is multiplied by the unit vector in the ith direction and the three components are added together
vectorially, we get
(4.17)

which is the differential statement of the law of conservation of momentum.


The combined momentum flux tensor ф is the sum of the convective momentum flux tensor ρvv and the molecular momentum flux
tensor π, and that the latter can be written as the sum of pδ and τ. When we insert ф = ρvv + pδ + τ into Eq. 4.17, we get the following
equation of motion:
(4.18)

In this equation p is a vector called the "gradient of (the scalar) p" sometimes written as "grad p." The symbol [ • τ] is a vector called
the "divergence of (the tensor) τ" and [ • ρvv] is a vector called the "divergence of (the dyadic product) ρvv."
THE EQUATIONS OF CHANGE IN TERMS OF THE SUBSTANTIAL DERIVATIVE

The observation of the concentration of fish in the River. Because fish swim around, the fish concentration will in
general be a function of position (x, y, z) and time (t).
𝜕
1. The Partial Time Derivative 𝜕𝑡 : -

Suppose we stand on a bridge and observe the concentration of fish just below us as a function of time. The time rate
𝜕𝑐
of change of the fish concentration at a fixed location. The result is ( 𝜕𝑡 ) x,y,z, the partial derivative of с with respect to

t, at constant x, y, and z.
𝑑
2. The Total Time Derivative 𝑑𝑡 : -

Now suppose that we jump into a motor boat and speed around on the river, sometimes going upstream, sometimes
downstream, and sometimes across the current. All the time we are observing fish concentration. At any instant, the
time rate of change of the observed fish concentration is

(4.19)

in which dx/dt, dy/dt, and dz/dt are the components of the velocity of the boat.
THE EQUATIONS OF CHANGE IN TERMS OF THE SUBSTANTIAL DERIVATIVE
3. The Substantial Time Derivative D/Dt: -
Next we climb into a canoe, and not feeling energetic, we just float along with the current, observing the fish
concentration. In this situation the velocity of the observer is the same as the velocity V of the stream, which has
components vx , vy , and vz . If at any instant we report the time rate of change of fish concentration, we are then giving

(4.20)

𝜕
The special operator D/Dt = 𝜕𝑡
+ v• is called the substantial derivative (meaning that the time rate of change is

reported as one moves with the "substance"). The terms material derivative, hydrodynamic derivative, and derivative
following the motion are also used.

For a function 𝑓, substantial derivative may be written as


𝐷𝑓 𝜕𝑓
= + 𝑣. 𝑓 (4.21)
𝐷𝑡 𝜕𝑡

𝐷𝑓 𝜕𝑓 𝜕𝑓 𝜕𝑓 𝜕𝑓
In Cartesian coordinates = + 𝑣𝑥 𝜕𝑥 + 𝑣𝑦 𝜕𝑦 + 𝑣𝑧 𝜕𝑧 where, 𝑓 is density, velocity, or energy.
𝐷𝑡 𝜕𝑡
More about THE SUBSTANTIAL DERIVATIVE (Gibbs notation)
The substantial derivative has a physical meaning: the rate of change of a quantity (mass, momentum, energy) as
experienced by an observer that is moving along with the flow. The observations made by a moving observer are
𝜕𝑓
affected by the stationary time-rate-of-change of the property ( 𝜕𝑡 ), but what is observed also depends on where the

observer goes as it floats along with the flow (v · 𝑓). If the flow takes the observer into a region where, for example,
the local energy is higher, then the observed amount of energy will be higher due to this change in location. The rate of
change from the point of view of an observer floating along with a flow appears naturally in the equations of change.

𝜕
Now we need to know how to convert equations expressed in terms of into equations written with D/Dt. For any
𝜕𝑡

scalar function f(x, y, z, t) we can do the following manipulations by using continuity equation in a flow of constant
density:

(4.22)
More about THE SUBSTANTIAL DERIVATIVE (Gibbs notation)
𝐷
The Equations of Change for Isothermal Systems in the 𝐷𝑡 -Form
Equation of Continuity at (4.7) can also be written as

(4.23)

and

Equation of Motion at (4.18) can also be written

as

(4.24)
𝐷
The Equations of Change for Isothermal Systems in the 𝐷𝑡 -Form
𝐷
The Equations of Change for Isothermal Systems in the 𝐷𝑡 -Form
𝐷
The Equations of Change for Isothermal Systems in the 𝐷𝑡 -Form
𝐷
The Equations of Change for Isothermal Systems in the 𝐷𝑡 -Form
𝐷
The Equations of Change for Isothermal Systems in the 𝐷𝑡 -Form
𝐷
The Equations of Change for Isothermal Systems in the 𝐷𝑡 -Form
𝐷
The Equations of Change for Isothermal Systems in the 𝐷𝑡 -Form
Equation of Motion
The most common simplifications of the equation of motion.
(i) For constant ρ and μ, insertion of the Newtonian expression for τ from Eq. 1.7 of Newton’s Equation into the
equation of motion leads to the very famous Navier-Stokes equation.

(4.24)
This equation can be written as
𝐷𝑉 𝐷𝑉
ρ 𝐷𝑡 = - p + μ 2V+ ρg or ρ 𝐷𝑡 = - +μ 2V
(4.25)
Equation of Motion
Equation of Motion
Equation of Motion
(ii) When viscous forces are neglected—that is μ = 0 or [ . τ] = 0, the equation of motion becomes

𝐷𝑉
ρ 𝐷𝑡 = - p + ρg (4.26)

It is known as the Euler equation for "inviscid" fluids or Ideal fluid.


Ex. Fluid flow away from the container or vessel walls. Such situation in metallurgy can be found in the
case of large furnaces like reverberatory furnaces and ladles.
FLOW THROUGH A CIRCULAR TUBE

Consider that steady-state, laminar flow of a fluid of constant density ρ and viscosity ρ in a vertical tube of length L
and radius R. The liquid flows downward under the influence of a pressure difference and gravity.
Assume the tube length be very large with respect to the tube radius, so that "end effects" will be unimportant
throughout most of the tube; that is, we can ignore the fact that at the tube entrance and exit the flow will not
necessarily be parallel to the tube wall.
We postulate that vz = vz (r), vr = 0, vθ = 0, and p = p(z). With these postulates it may be seen from Table B.I that the
𝑑𝑣
only nonvanishing components of τ are τrz = τzr = - μ 𝑑𝑟𝑧
Cylindrical shell of fluid over which the z-momentum balance is made for axial flow in a circular tube
The momentum-flux distribution and velocity distribution for the downward flow in a circular tube.
We now summarize all the assumptions that were made in obtaining the Hagen Poiseuille equation.

(a) The flow is laminar; that is, Re must be less than about 2100.

(b) The density is constant ("incompressible flow").

(c) The flow is "steady" (i.e., it does not change with time).

(d) The fluid is Newtonian

(e) End effects are neglected. Actually an "entrance length," after the tube entrance, of the order of Le = 0.035D Re,

is needed for the buildup to the parabolic profile.

(f) The fluid behaves as a continuum

(g) There is no slip at the wall, so that B.C. 2 is valid;

You might also like