You are on page 1of 41

Fluid flow in process units

1 Introduction ......................................................................................................................................... 2
2 Mechanical energy balances ................................................................................................................ 2
2.1 Friction ......................................................................................................................................... 3
2.2 Friction in empty pipes ................................................................................................................. 3
2.3 Local friction losses....................................................................................................................... 5
2.4 System curve ................................................................................................................................ 5
2.5 Pump curve .................................................................................................................................. 5
2.6 Pump cavitation ........................................................................................................................... 6
2.7 Fans, Blowers and Compressors.................................................................................................... 6
2.8 Flow in porous media ................................................................................................................... 7
3 Momentum balances ........................................................................................................................... 8
3.1 Mechanical energy vs. momentum balances................................................................................. 8
3.2 Newton's viscosity law.................................................................................................................. 9
3.3 Laminar flow in a pipe .................................................................................................................11
3.4 General momentum balances......................................................................................................14
4 Non-Newtonian fluids .........................................................................................................................15
5 Turbulence ..........................................................................................................................................17
5.1 Turbulent energy spectrum .........................................................................................................19
5.2 Turbulence modeling...................................................................................................................20
5.3 Turbulence near walls..................................................................................................................21
6 Multiphase flow ..................................................................................................................................22
6.1 Flow regime maps and pressure drop ..........................................................................................23
6.2 Interphase forces.........................................................................................................................24
6.3 Fluidization..................................................................................................................................27
6.4 Two-phase flow in porous media .................................................................................................28
6.5 Capillary pressure and surface wetting ........................................................................................29
7 Mixing.................................................................................................................................................30
7.1 Mixing in stirred tanks .................................................................................................................30
7.2 Multiphase mixing in stirred tanks ...............................................................................................32
7.3 Static mixers ................................................................................................................................33
8 Computational fluid dynamics (CFD)....................................................................................................34
8.1 Evaluation of the grid ..................................................................................................................35

1
8.2 Modeling of equipment with moving parts ..................................................................................37
8.3 Computational modeling of multiphase systems..........................................................................37
8.4 Modeling of moving fluid interfaces ............................................................................................38
9 Flow measurement .............................................................................................................................39
10 Symbols ..............................................................................................................................................40

1 Introduction
This course gives the basics for fluid flow on master level studies. It first reviews basics of pump/piping system
design using mechanical energy balances. This is typically a BSc level topic, but due to its importance in
chemical engineering design, it is reviewed on this course. Most of the fluid flow problems in practice are
solved with this approach appended with a general engineering judgement. In addition to that, this course
gives overview on various fluid flow related issues, such as non-Newtonian flow, flow regime maps in two-
phase flows, mixing in stirred tanks and static mixers, and flow in porous media. In addition, turbulence and
computational fluid dynamics are briefly introduced.

These lecture notes follow the same themes as the course, but are not necessarily presented in precisely the
same order. This lecture note is not either designed to be fully stand-alone; some things are explained very
briefly here and should be explained in more detail during the lectures. You are also by no means restricted
to find information in this lecture note; this is intended to give an overview to help following the lectures and
to understand other information sources (books, articles, web pages, presentations, discussions with
colleagues etc.) better.

2 Mechanical energy balances


General energy balance can be considered to be “always valid”, provided all the relevant source terms are
properly taken into account. The other balances always needed are the material balances (total mass or
chemical component masses or moles). In many fluid flow related cases also momentum balances are
needed, although this is true for mainly in such situations, where precise flow patterns are needed. In many
cases this is not necessary. Equivalence of mass and energy as treated by high energy physics is not either
needed in chemical engineering applications; mass and energy can be considered completely independent
with their own balances.

General energy balance is typically divided in two parts: mechanical energy and thermal energy. The reason
to do so is that in different applications one or the other form is only needed. There is naturally
transformation of mechanical energy to thermal energy (due to friction losses) or in some cases
transformation of thermal energy to mechanical energy (in heat engines), but these transformations are
typically taken into account by specific source terms.

Mechanical energy balance can be written in the following way:

2

θγp Wp < ∋p b , p a ( ∗ θg ∋z b , z a ( ∗ 1 2 θ  a v b2 ,  b v a2 ∗ θh f ( (1.)

Where the first term on the left is (shaft) work done by the pump, the first term on the right is pressure
difference between the end and the source point, the second term is hydrostatic pressure due to height
difference between end and source points, the third is pressure change due to velocity (also called dynamic
pressure), and the last term on the right is due to friction losses.

This equation without the shaft work by the pump and the friction loss term is called Bernoulli equation,
which is then a mechanical energy balance with some assumptions (no work or friction). It is applicable to
some specific cases, although the general form, which is sometimes called Extended Bernoulli equation, is
much more general and should be used in most cases.

Out of the various terms, the dynamic pressure is very small compared to the others in almost all cases. For
constant flow cross-sectional area and incompressible flow, it is precisely zero. Relative difference of the
other terms depend on the case; for horizontal pipes the hydrostatic pressure difference is zero, and for flow
between two vessels open to the same gas space (atmosphere etc.), the pressure difference term is zero.

Note that the term "pressure drop" is often used to estimate mechanical energy loss when it dissipates to
heat. This is not necessarily the same as pressure difference, as the latter may include hydrostatic head.
Pressure drop in a straight vertical pipe is the same irrespective of flow direction (in one-phase flow), but
pressure difference between beginning and end of the pipe depend on the flow direction.

2.1 Friction
The friction term is typically divided into friction caused by empty pipe (viscous dissipation at pipe walls) and
local friction losses:

∑ ΧL
2
⌡v
θh f < θ ω ∗  ψi  (2.)
 D  2

The first term in parenthesis describes friction caused by empty pipe (it is proportional to the pipe length ΧL).
The proportionality coefficient ω is called the Darcy friction factor. The second term in the parenthesis is a
sum of all the local friction losses caused by pipe bends, valves, instrumentation etc. Even heat exchangers,
reactors and other process equipment can be taken into account in a similar way. The formulation is written
so that dependency on velocity is visible. For a fully turbulent flow, the friction factors in the parenthesis are
relatively constant (independent on velocity), so that pressure drop will be proportional to the velocity
squared.

2.2 Friction in empty pipes


Friction (Darcy friction factor) in empty pipes can be estimated from the Moody diagram:

3
As can be seen, for a fully turbulent regime, the friction factor becomes constant after sufficiently high flow
rate. This depends also on the surface roughness of the pipes, which is a property of the material and
manufacturing method.

As these diagrams are not very convenient for normal design calculations which require iterative solutions
to various variables, equations are typically used. The fully turbulent regime friction can be well estimated
using the Colebrook equation:

1 ∑ k 2.51 ⌡
< ,2.0 log ∗  (3.)
ω  3.7D Re ω 

As can be seen, the friction factor ω is implicit and can be solved only iteratively. There are various explicit
approximations that are of sufficient accuracy for most engineering purposes, for example:

1 ∑ k 5.02 ∑ k / D 13 ⌡ ⌡
< ,2.0 log , log ∗   (4.)
ω  3.7D Re  3.7 Re  

For laminar pipe flow, there is an analytical solution to the friction term. This will be discussed later with the
momentum balances.

4
2.3 Local friction losses
All additional disturbances to well-developed pipe flow cause additional pressure losses. These are case
dependent, and are calculated as local frictional losses or resistance coefficients. In principle it is rather
straightforward to look at appropriate contributions and add them for the summation term in equation 2. It
does not matter in which order the local losses are in the direction of flow, and how long empty pipes are in
between. Resistance coefficients may depend on the pipe diameter. Sometimes pipe entry from a vessel is
included as local friction loss; sometimes it is included in the dynamic pressure discussed earlier. Typically
this term is small, but should be considered consistently. The largest local pressure drop values are typically
caused by valves, where significant (typically at least tens of kPa) pressure drop is required even for a fully
open valve due to controllability reasons. If the pressure drop is higher somewhere else, changing the valve
position would not make a significant modification to the flow rate, which is not desirable situation. If whole
equipment (heat exchangers etc.) are included as local pressure drops, these could be of comparable to
control valve pressure drop values, but preferably not higher.

2.4 System curve


When pressure change is plotted against flow rate, the so called system curve is obtained. This curve
describes pressure changes in the whole piping system as a function of flow rate. For turbulent flow (as is
typical in chemical engineering industrial applications), the curve follows a second order polynomial form Χp
= a + bv2, where a contains pressure difference and hydrostatic head (see mechanical energy balance,
equation 1), and b contains possible dynamic pressure effect and most notably the friction terms. Typically
the system curve is plotted in the same figure together with the pump curve shown in the next chapter.

2.5 Pump curve


Pressure increase in the pump as a function of flow rate follows a characteristic shape for the selected pump
type. The most typical pump type in chemical industries is centrifugal pump. Hypothetical centrifugal pump
and two piping curves are drawn in the same figure here:

5
Y-axis is often plotted in terms of height (head), so that its dimension is meters (or feet) instead of pressure.
Pressure can be changed to head by dividing by fluid density and gravitational acceleration. The pump curve
(black) crosses system curve (red or blue) at the operating point. At this point the left and right hand sides of
the mechanical energy balance are equal.

The two system curves in the figure could represent situations with different openings of a control valve.
Valve is originally fully open (flow v1 and pressure difference p1), and then partially closed (v2 and p2).
Naturally a fully closed valve would lead to infinite local resistance coefficient with zero flow rate. In that
situation the pressure next to the pump is at maximum, corresponding to the point where pump curve
crosses y-axis. The piping system must be designed to mechanically tolerate at least this pressure and some
safety margin.

2.6 Pump cavitation


In some cases there is a danger that the pumped liquid will form bubbles inside the pump. This can happen
if the liquid is relatively hot compared to its bubble point, i.e. close to boiling. As there is some pressure drop
in the pump feed piping, the pressure inside the pump may be lower than the bubble point pressure, leading
to bubble formation. These bubbles collapse again as the pressure in the pump is increased. This process
causes noise and harms the pump. It is important to prevent this by designing the pump feed piping system
in an appropriate way. The most common way is to position the pump below the feed vessel, so that the
pressure in the pump is higher than in the feed vessel due to hydrostatic head.

It is important to prevent also gas bubbles from the feed vessel to enter the pump, at least continuously. This
can be done by proper design of the vessel itself. This topic may seem trivial, but is essential in all chemical
engineering applications; control of the phases in the sense that phases are either in good contact (mixed
well enough for the purpose but not more) or separated and allowed to flow at desired destinations (e.g.
settling vessels).

2.7 Fans, Blowers and Compressors


These are equipment that move and compress gases. Fans discharge large volumes of gas into open spaces
or large ducts. They generate low pressure increases, in the order of 0.04 atm. Blowers are high-speed rotary
devices that develop maximum pressure up to 2 atm. Compressors can compress gases into very high
pressures using several stages.

In calculations, fans can be treated as pumps as the density of gas is not changing. The main principal
difference between pumps and compressors is that compressors are pumping gases, which are compressible.
As with pumps, there are many different compressor types. Compressors can be crudely divided into positive
displacement and dynamic compressors. Positive displacement compressors can further be divided into
rotary and reciprocating compressors. Dynamic compressors are either axial or centrifugal type.

The most important design variable for compressors is perhaps the required compressing power. Typically
compressors are assumed either isentropic (zero entropy generation, no cooling) or polytropic (some cooling,
can be used for large compressors with cooling that is not perfect). For small and well cooled compressors,
isothermal assumption can also be used, but not needed so often in practice. Since it is difficult to increase
pressure very much in a single compressor stage (except for reciprocating compressors), they are often built
as multi-stage apparatus.

Power requirements (J/mol) for single-stage compressors can be estimated from:

6
nRT1 ∑ p ⌡ ∋n ,1( / n 
Wcompr <  2  , 1 (5.)
n ,1  p1  

This gives the ideal compression work, which needs to be corrected by dividing with compressor efficiency
γcompr (e.g. 0.7). In order to estimate power (W), this must be multiplied by the molar flow of gas (mol/s).

For isentropic compressors, n is the ratio of heat capacities, n=k=Cp/Cv (for air this is approximately 1.4). This
can be also used for reciprocating compressors. For polytropic compression of ideal gases, n can be estimated
with the following correlation:

1
n< (6.)
k ,1
1,
kγcompr

In practical compressor calculations, especially at high pressures, it is important to specify also


thermodynamics (equation of state) properly, as gases may be far from ideal at high pressures. In these cases,
correlation for n is more complicated than above.

2.8 Flow in porous media


In many cases, solid materials are either treated in chemical engineering applications or used as catalyst or
other processing media. For example biomass treatment, filtration, packed bed reactors etc. are very
common. This chapter considers only situations where the solid phase is stagnant, i.e. solid particles are not
moving. Fluidization and other applications where solids are moving are treated separately. Only one fluid
phase is also considered here, multiphase flow is discussed in other chapters.

Typical to these situations is that the precise structure of the porous media is not known, but average
properties are known instead. The known properties typically include porosity (empty space available for the
fluid to flow) and particle size and shape. The latter can be tedious in case of wide distributions both in size
and shape, as is often the case in biomaterial treatment.

A layer of porous material in the fluid flow path can be considered as a local friction loss. However, in most
cases the unit consisting the porous bed is designed separately during the preliminary design phase, so that
its pressure drop is evaluated separately from the other pump and piping system. Later on, the porous bed
can be accounted in the system curve discussed earlier.

The most classical pressure drop equation for flow in porous media is the Ergun equation:

λ∋1 , δ ( θ∋1 , δ ( 2
2
Χp
< Eλ j ∗ Eθ j (7.)
L d pδ
2 3
d p δ3

7
Here the parameters Eλ and Eθ are empirical constants. In the original paper by Ergun, Eλ =150 and Eθ = 1.75
were recommended, but later on based on more experimental data, more conservative values of Eλ =180
and Eθ = 1.8 are proposed. j is the superficial velocity, i.e. linear average velocity of the fluid if it would flow
through the empty vessel without solids. If particles are not completely similar (spherical with constant
diameter), an effective diameter needs to be used.

Another way of expressing pressure drop – flow rate relationship in porous media is to use permeability
concept. It is typically used for relatively slow flow rates (laminar flow), and in materials that are not
composed of separate particles, but materials with natural pores in which fluid may flow. Permeability can
be defined by the Darcy's law:

kA Χp
Q< (8.)
λ ΧL

where k is permeability and A is cross-sectional area. If the relationship j = Q/A is used, it is easy to see that
permeability for laminar flow in porous material can also be predicted with the Ergun's equation.

3 Momentum balances
In some cases, the pressure drop alone is not sufficient, but we want to know more about the flow patterns.
In these cases mechanical energy balance alone is not sufficient. Momentum balances are in many respects
more demanding to formulate properly and especially to solve. One fundamental challenge comes from
momentum transfer due to viscous forces. Mass transfer (diffusion) and heat transfer (thermal conduction)
can often be analyzed with a one-dimensional model: heat is conducted from hot to cold and diffusion
transports molecules from high concentration to low (statistically). Instead, momentum is transferred
perpendicular to the flow direction. For example linear momentum in x -direction is transferred by viscous
forces in y- and z –directions. Flow therefore "grabs" fluid elements next to it to move in the same direction.
Near the walls the fluid is stagnant (so called no slip condition), and this process transports linear momentum
from the main flow towards the stagnant walls. In this process, part of the fluid mechanical energy is lost due
to dissipation to heat. The correlations in the previous chapter estimate this in some well-defined systems or
based on empirical correlations. If these are not available or the precise flow pattern needs to be known for
design purposes, linear momentum balances need to be used.

3.1 Mechanical energy vs. momentum balances


Typically mechanical energy balance, or energy balance in general, seems to be easier to comprehend than
linear momentum balances. This may be due to their wider use in classical problem solving in schools. Here
some simple relations are first discussed to clarify the two concepts.

The dimension of energy (extensive property for which a conservation equation or balance can be written) is
Joule, so that time dependent energy balance has terms that are in the dimension of J/s, or W. Sometimes
energy balances are called therefore power balances, although the conserved property is energy. For fluid
flow applications, power is simply a product of pressure drop and volumetric flow rate:

8
P < QΧp (9.)

Power also equals to force times velocity.

Linear momentum is mass times velocity, dimension (kg m/s). Time dependent momentum balance contain
terms that are of the same dimension than rate of change of linear momentum (kg m/s 2). This is the same as
force (N). Actually many mechanics problems are classically solved with "force balances", which can be
understood as time independent linear momentum balances, where the system is not accelerating. Then
various forces compensate each other. The conserved extensive property, however, is linear momentum, not
force. Force and pressure applied to a surface are related as

F
p< (10.)
A

This is one of the terms appearing in the linear momentum balances for fluid flow problems as shown later.

In addition to the linear momentum, also angular momentum is conserved. However, angular momentum
balances are not typically separately formulated in chemical engineering fluid flow problems and not
discussed further here. For differential momentum balances, symmetry of the stress tensor implies
conservation of angular momentum on small scales.

3.2 Newton's viscosity law


Let's start analyzing linear momentum balances by reviewing Newton's viscosity law. Perhaps the most
classical way of explaining this is to consider liquid layer between two parallel plates that are very large. Large
plates are assumed just to neglect any boundary effects. The plates are initially at rest, and the fluid stagnant.
At some time moment, one of the plates is being pulled with a force F:

The other plate remains stagnant. The fluid elements just next to the moving plate move with the same
velocity as the plate, and fluid elements just next to the stagnant plate remain stagnant. After a while, a
velocity profile is formed in the fluid between the plates:

9
The force divided by plate area is called shear stress:

F
<σ (11.)
A

It is experimentally found that the shear stress is proportional to the shear rate (velocity gradient):

dv x
σ yx < ,λ (12.)
dy

This proportionality is called Newton's viscosity law. Here subscript yx for the shear stress refer to x-
directional linear momentum being transferred into y-direction. In principle it is transferred also to z-
direction (the third dimension not shown in the figure), but as the system is symmetric in that direction in
this example (constant velocity), the net transfer in that direction is zero and can be excluded from this
analysis. Minus sign refers to the fact that momentum is transferred to the direction of lower velocity.

Newton’s law for viscosity could be compared to Fick’s law for diffusion and Fourier’s law for heat conduction:

dc A dT
Diffusion (Fick): J A < , D AB Heat conduction (Fourier): q < ,κ
dy dy

It is clear that these laws are mathematically quite similar including flux, a proportionality coefficient, and
assumed linear dependency on gradient related to the modeled variable. Therefore, the solutions are also in
many respects similar. The challenge in fluid flow is that there are always more than one spatial coordinate
involved, as linear momentum in x direction (in this case) is transported in y –direction. For diffusion and
heat conduction, often one-dimensional treatment is sufficient.

The proportionality coefficient in the Newton's viscosity law is viscosity. For many fluids it does not depend
on the shear rate or time, i.e. is practically constant. These are called Newtonian fluids. If the proportionality
coefficient depends on the shear rate or time, the fluid is called non-Newtonian. In all cases, viscosity is a
function of temperature.

10
Rheology is a study discussing properties of flowing substances, and characterization of non-Newtonian fluids
is an essential part of it. These will be discussed later.

In general three-dimensional cases, collection of each shear stress component is called stress tensor.

3.3 Laminar flow in a pipe


A classical application example of Newton's viscosity law is a well-developed laminar flow in a circular pipe.
Well-developed here means, that any entry effects at the pipe feed have been disappeared from the flow
profile. We need to formulate a linear momentum balance for a control volume shown in the next figure:

The control volume is between the two cylindrical regions at arbitrary distance r from the pipe center axis.
The control volume is of thickness Χr and of length ΧL. Flow is assumed to be from left to right.

Linear momentum flow in to the control volume through the control volume boundary on the left (ring at the
left end of the cylinder) and linear momentum flow out from the right end are

in < ∋2 ο r Χr v z (∋θ v z ( z < 0

11
out < ∋2 ο r Χr v z (∋θ v z ( z < L

Since the cross-sectional area of the pipe is constant and the fluid here assumed incompressible (constant
density), these two terms are equal and cancel each other out from the balance.

Another source term in the linear momentum balance is due to pressure at the same two ends. Force equals
pressure times area, so the pressure terms are:

in < ∋2 ο r Χr (p 0
out < ∋2 ο r Χr (p L

Now pressure at the inlet (p0) is not the same as pressure at the outlet (pL). This pressure difference is the
driving force for the fluid to move, and the related energy is dissipated due to viscous forces. In order to
describe these, we need the shear stress definition:

in < ∋2 ο r ΧLσ rz ( r

out < ∋2 ο rΧ Lσ rz ( r ∗ dr

where the cylindrical part area is multiplied by the shear stress at the corresponding location (r and r+Χr).
The linear momentum balance thus reads

0 < ∋2 ο r ΧLσrz ( r ∗ 2 ο r Χrp0 , ∋2 ο r ΧLσrz ( r ∗ Χr , 2 ο r ΧrpL (13.)

In order to close the equation, we need to use the Newton's viscosity law for the shear stress. After inserting
it and allowing Χr to approach differentially small thickness, we end up with the following equation:

d∑ dv ⌡ ∑ p , p L ⌡
, rλ  <  0 r (14.)
dr  dr   ΧL 

12
This second order ordinary differential equation can be solved with two boundary conditions for the variable
depending on r (velocity). Note that pressure drop per length is assumed constant and thus the equation
here is not a partial differential equation.

The two conditions are no slip at the wall (discussed earlier), and symmetry at the center. The latter boundary
condition results from the fact that for the shear stress to be defined at the center, the shear rate needs to
be smooth (i.e. zero). The solution to this is

∑ p 0 , p L ⌡ 2 ∑ ∑ r ⌡ ⌡
2
∑ p0 , p L ⌡ 2
v∋r ( < , ∋
 r , R < 
2
(  R 1 ,  
   R  
(15.)
 4 λ ΧL   4 λ ΧL

This gives velocity at any distance from the center of the pipe. It can be seen, that the velocity is a second
order polynomial form. Maximum velocity is at the center and minimum (zero) near the walls.

We are often interested in the average velocity, as it is related to the volumetric flow rate. Unless otherwise
specified, it is always the average velocity that is referred to when fluid flow rate in pipes is expressed in m/s.
Average velocity can be obtained by integrating the local velocity profile:

R
Q < 〉 v∋r (2οrdr < vοR 2 (16.)
0

The result is

Q p 0 , p L 2 v max
v< < R < (17.)
οR 2
8λL 2

Here it can be seen that the maximum velocity at the centerline is twice the average velocity. The previous
equation is often written in the form explicit to pressure drop:

8λΧL 32λΧL
Χp < v < v (18.)
R2 D2

13
This equation is called Hagen-Poiseuille law. Often it is also given in terms of volumetric flow instead of linear
flow velocity. Note that this is to some extent similar than the laminar contribution of the Ergun equation,
but restricted to circular straight pipes.

When this is compared to the definition of Darcy friction factor (part of equation 2), we see that

64λ 64
ω< <
θDv Re (19.)

This is an analytical solution for the Darcy friction factor for purely laminar flow in circular pipes. It can be
seen also in the Moody diagram for laminar flow velocities (Reynolds number below approximately 2000).
This is one of the few cases where momentum balances can be solved analytically. In most cases a numerical
solution is needed. There are well established methods and software for this. This field of fluid flow study is
called Computational Fluid Dynamics (CFD), and is briefly introduced later.

3.4 General momentum balances


Unfortunately the geometry is not always as simple as a circular pipe. If we want to calculate flow patterns
in more complicated situations, we need general momentum balances. There will be three of them, one for
each spatial coordinate. General differential momentum balances can be formulated based on the following
control volume sketch:

When all the directions are taken into account and all the relevant source terms, the following equations
result in:

∝ ∋θU i ( ∝ ∋θU i U j ( ∝ ∋σ (
∗ < ∗ Sε (20.)
∝t ∝x j ∝x j

14
Note that this is the same form as a general transport equation, e.g. differential material or energy balance.
Here the transported property is linear momentum divided by volume, θU. Note that in this context symbol
U is used for velocity (for historical reasons).

Here so called Einstein notation is used, so there are actually three separate equations for the linear
momentum; subscript i here refers to the separate equation whereas subscript j refers to repeated indices
(summation).

When Newton's law is used for shear stress tensor σ, and gravity and pressure gradients are observed to be
a source term, we end up with

∝ ∋U i ( ∝ ∋U i ( ∝p ∝ ∑ ∝U i ∝U j ⌡
θ ∗ θU j < , i ∗λ ∗ ∗ θg i
∝t ∝x j ∝x i ∝x j  ∝x j ∝x i  (21.)

In the previous equation, also differential material balance is used, and viscosity is assumed constant.
Detailed derivation of these equations can be found in the fluid flow literature.

The previous equations, along with the differential material balance, are called Navier-Stokes (N-S)
equations. The first term on the left describes rate of change of linear momentum, the second term
convection of momentum, the first term on the right is pressure gradient source term, the second viscous
dissipation and the last term source term due to gravity. In this course, we are not extensively working with
the N-S equations by hand, but later on discuss how it is typically solved with proper algorithms.

4 Non-Newtonian fluids
For Newtonian fluids, the shear stress is linearly proportional to the shear rate, and this relation does not
depend on time. For non-Newtonian fluids, the dependency is more complicated. Non-Newtonian fluids can
be characterized based on their shear stress – shear rate dependency, shown in the following figure

15
Typically, pure fluids with low molecular weights have close to Newtonian behavior, e.g. water or light
hydrocarbons. Pseudoplastic fluids are also common, e.g. syrup, pulp in water, and many fermentation
broths are pseudoplastic. Cornstarch and some colloidial systems are dilatant. Drilling mud, toothpaste,
mustard are some examples of Bingham –type behavior where the fluid has a yield stress before it starts to
flow at all.

Viscosity in these systems can be characterized by shear rate dependent apparent viscosity. It is the slope of
such curve plotted through measured shear stress vs. shear rate point in the figure above and origin.

Further complication is that in some cases apparent viscosity is time dependent. This can occur e.g. in
situations where the fluid consists of long chain macromolecules, which take time to orient in the direction
of the flow, or in cases where other reversible structural changes happen in the fluid once subjected to shear.

There are several models for non-Newtonian fluid viscosities. Perhaps the simplest is the power law:

n
∑ dv ⌡
σ < m   (22.)
 dy 

which leads to the following apparent viscosity model:

16
n ,1
∑ dv ⌡
λ < m   (23.)
 dy 

Another, more versatile is Carreau model, which is suitable for a wide range of shear rates, but requires also
experimental apparent viscosity measurements over a very wide range of shear rates in order to identify the
unknown four parameters.

∋ n ,1( / 2
λ , λ ⁄ ∑ ∑ dv ⌡ ⌡
2

< 1∗ κ 
λ 0 , λ ⁄   dy  
(24.)

5 Turbulence
When the flow rate is increased, the destabilizing inertial forces become larger than the stabilizing viscous
forces, and the flow becomes unstable. The threshold for this can be estimated with the Reynolds number:
θvD
Re < . There is no precise value for this transformation, and it depends also on the definition of the
λ
characteristic length used in the Reynolds number. There is also a transitional regime between the laminar
and fully turbulent flow, where the flow is not stable, but there are not yet such statistical variations in the
local flow velocities that are characteristic to fully turbulent flow. Depending on the case, above Reynolds
number around 4000 – 10 000, the flow can be assumed to be fully turbulent. Interestingly, the Reynolds
number appears also naturally when the Navier-Stokes equations are made dimensionless. This underlines
the fundamental nature of the Reynolds number in the field of fluid dynamics.

Although the N-S equations basically can be used to model precisely all possible fluid dynamics problems,
turbulent fluctuations are typically too rapid and small for carrying direct numerical simulation in practice
(more about this later). Therefore the turbulent flow is typically modeled with the so called Reynolds
decomposition, where the instantaneous velocity is split into time-averaged and fluctuating part:

U i < U i ∗ u i' (25.)

This division is illustrated in the next figure.

17
Note that the time-averaged part U i does not imply a steady-state assumption, but the changes in the
average flows are just assumed to be much slower than turbulent fluctuations.

When the Reynolds decomposition is applied to N-S equations, a time-averaged form is obtained:

∝∋U i ( ∝∋U i ( ∝p ∝ ∑ ∑ ∝ U i ∝ U j ⌡ ⌡
θ ∗ θU j <, i ∗ λ ∗ , θu 'i u 'j  ∗ θg i
∝x i ∝x j   ∝x j ∝x i   (26.)
∝t ∝x j 

Here it can be seen, that the equation is otherwise similar than the original N-S, but there is an extra term

σij' < θu i' u 'j (27.)

This term is called Reynolds stress tensor (Reynolds stresses), and it describes turbulent momentum
transport. Note that it appears in the same part of the equation as viscous dissipation. It takes into account
a process where eddies (small scale liquid vortices) carry linear momentum to slower moving regimes of the
fluid in a similar fashion as viscous forces do (see discussion related to the Newton's viscosity law). As a result,
it gives additional friction-like effect to the flow.

As small scale fluctuations are not known (unless the full non-decomposed N-S equation is solved, which we
try to avoid in this whole process), the Reynolds stress term needs to be modeled somehow. There are
various attempts to do this especially as related to the computational fluid dynamics. These models are one
example of so called “closure models”, which are in practice often needed in physical modeling. These are
discussed later.

18
5.1 Turbulent energy spectrum
Various eddies in the flow can be characterized based on their size and the typical kinetic energy they contain.
A classical scheme to illustrate this is the following:

This is often called Kolmogorov energy spectrum or energy cascade. Largest eddies that are created from the
main flow are on the left and smallest dissipating kinetic energy to heat are on the right. The largest eddies
move faster and contain more energy, and the small eddies contain less energy (per eddy). However, there
are much more small eddies, so on average the total energy in various eddy sizes is rather constant until the
smallest eddies which are no longer broken into smaller ones but the energy is dissipated to heat. Smallest
eddies are characterized by the so called Kolmogorov length and time scales:
1/ 4
∑ δθ3 ⌡
Length scale γ <  3  (28.)
λ 
1/ 4
∑λ⌡
Time scale σ γ <   (29.)
 δθ 

Where δ is the local energy dissipation to heat (W/kg). It is a very useful variable also when total energy input
to a system is evaluated. For example in stirred tanks, average δ value is the total mixing power per fluid
mass. It is a quick and convenient value to evaluate how strong mixing is used. Typically its values are around
0.01-10 W/kg in liquid mixing. This energy is then dissipated all around the mixing tank, but not equally. Near
the impeller the dissipation is much more intense, and based on Kolmogorov time and length scales, also
eddies are smaller and more rapid.

19
5.2 Turbulence modeling
Turbulence modeling typically refers to various efforts to model Reynolds stress tensor empirically. This is
needed as there is no analytical solution to the turbulent flow problem. The most typical approaches are
based on the so called Boussinesq approximation. Based on that assumption, turbulent contribution is added
to the viscosity:

λ ↑ λ molecular ∗ λ T (30.)

Note that the turbulent viscosity λT is a property of the flow, not of the fluid. When turbulent viscosity
approach is used, the time averaged N-S equations appear essentially the same as the original ones, but
viscosity replaced with the combination of molecular and turbulent contributions. In fully developed
turbulent flows (high Reynolds numbers), the turbulent viscosity is much higher than the molecular one.

Turbulence models are often characterized by the number of additional transport equations needed to model
the Reynolds stress tensor, or turbulent viscosity if that is assumed. Some of them are listed in the following
table:

In practice, 0 and 1 equation models are not used. Two-equation models are currently perhaps the most
often used, as they are a convenient compromise between computational burden and accuracy. Perhaps the
most typical is the so called k-δ model (at least for chemical engineering applications), where one transport
equation (new differential balance equation for a modeled property) is formulated for kinetic energy of the
turbulent fluctuations

20
1 3 '2
k <  ui (31.)
2 i <1

And one for the turbulent kinetic energy dissipation rate δ discussed earlier. Formally the latter is not a
conserved property, but nevertheless a transport equation can be formulated it with apparently good
predictive capabilities. There are specific source terms (production and disappearance) for k and δ as well as
convection and other terms. These equations can model so called history effects, where turbulent
fluctuations move with the main flow while the eddies are dissipating fluctuation energy. Eddy viscosity can
be directly then calculated with the two additional variables. Note that the total set of equations to model
three-dimensional turbulent flow would in this case be six: one material balance, three linear momentum
balances (one for each direction), one equation for k and one for δ. Turbulence is thus assumed isotropic as
there is no preferred direction in the modeled eddies.

In more rigorous turbulence models, such as Reynolds stress model and large-eddy simulations (LES), the
directional character of larger eddies are taken into account. In Reynolds stress model, this is done by using
several additional transport equations, and in LES by modeling numerically all except the smallest eddies.

The most rigorous model is Direct numerical simulation (DNS), where no turbulence model is needed. The
numerical solution needs to be so accurate that all the smallest eddies are resolved. This requires a
computational grid where the grid size is smaller than the Kolmogorov length scale and time dependent
solution where the time step is shorter than the Kolmogorov time scale. Even then the solution is chaotic in
nature; we cannot say for sure that at a certain point in space at a given time the flow is precisely as predicted,
the result is just one possible realization of the flow.

5.3 Turbulence near walls


Near solid walls, and also near fluid interfaces, turbulence tends to decay. Large eddies simply cannot exist
very close to the wall for physical reasons. Near the walls velocity gradients can be very steep, and molecular
viscosity becomes significant even for fully turbulent flow. Near phase boundaries (solid-fluid or fluid-fluid)
are of high importance in chemical engineering, as profiles close to the walls or interfaces determine mass
and heat transfer rates. The following figure illustrates the processes near boundaries, and revises the
classical film theory which can be applied both to heat and mass transfer.

21
In the film model, there is a hypothetical film thickness next to a wall or phase boundary. This thickness is
determined so that the gradient next to the wall is the same as in the “true” profile, so that mass or heat
transfer rate due to diffusion or thermal conduction is the same in both models. The whole film is assumed
laminar, and its thickness is typically not needed separately, but combined with diffusion coefficient or
thermal conductivity into mass or heat transfer coefficient.

In fluid flow boundary layer, also a laminar film is assumed. However, this is true laminar layer next to the
wall, and its thickness is not the same as in the film model (it is thinner). Next to it, a so called “universal
velocity profile” is assumed, where decay of turbulence and other flow properties can be modeled with
correlation approach.

It is well known that some turbulence models do not predict well fluid flow in the boundary layer. Most
notably k-δ model is not good at this region. One of the reasons is that near the wall the turbulence is no
longer isotropic, i.e. flow fluctuations are not similar in directions perpendicular to the wall and adjacent to
the wall. In order to help in this, so called wall functions are used along with such turbulence models, so that
the flow near the walls is predicted with empirical wall functions and the turbulence model is applied outside
this region.

6 Multiphase flow
In many chemical engineering applications, multiple phases flow simultaneously. This could be just gas and
liquid phases flowing together in pipes, flow in multiphase reactors, such as gassed stirred tanks, bubble
columns, or trickle beds, or in most separation processes, where the process performance is based on good
contact between two immiscible phases and distribution of the separated components between these
phases, such as in distillation or extraction columns. The first, and perhaps even the most important, skill for
a chemical engineer in these cases is to be able to visualize the situation. This is required for sensible design
of the units so that the phases flow as designed; either providing good contact for mass transfer, or provide
good separation of the phases e.g. in settlers. For good contact, typically energy input is needed to disperse

22
one phase into another. In many cases, this energy is provided by pressure drop (power equals pressure drop
times volumetric flow), and in some cases it is provided mechanically, such as in stirred tanks. For proper
separation of the phases, the simplest approach is to allow slow enough flow so that dispersed phase bubbles
or droplets have sufficient time to settle. A classical design approach for this kind of operation would be
calculating fluid velocity for self-venting flow, where liquid velocity downwards should not be faster than the
assumed bubble rise velocity, if the bubbles need to be separated into a gas space on top of the liquid. In
these cases, the cross-sectional area available for the liquid (or dispersion) flow should be large enough so
that the velocity is below the rise velocity of bubbles. In these cases, design requires that a threshold bubble
size is assumed. Smaller than this threshold sized bubbles will not be properly separated since they rise
slower than the liquid velocity, but larger ones should separate. Another example is gas-liquid flow in
vertically placed reactors: in downward flow arrangement gas accumulation in the reactor is relatively large
and cannot be avoided, whereas in upwards flow arrangement gas escapes more rapidly from the reactor
space. Which one is the desired operation, depends on the case. There are several such examples where
rather simple common sense helps designing units, at least giving qualitative reasoning for the design. When
this common sense is combined with quantitative models capable of predicting the flow, proper design can
be achieved.

6.1 Flow regime maps and pressure drop


Two (or more) phases may flow in completely different flow patterns, or textures, depending on the
conditions. The most important parameter is probably flow rate of each of the phases, flow direction, and
the cross-sectional area available for the flow. Also physical properties, such as liquid phase viscosity and
surface tension, may affect flow pattern. Based on the flow rates, various patterns can be found from
available diagrams. Note that the limits are not necessarily precise. In the figure below, schematic flow
patterns are shown on the left, and one available flow regime map on the right for horizontal flow.

One problem in these maps is that in many cases they are based on experimental data with rather narrow
validity range. Most of the data is measured by using air as gas and water as liquid, and in pipes with
diameters typical for laboratory equipment. These are sometimes corrected for varying properties with
separate correction factors. It is, however, clear that for example slug occurs much more easily in very small
pipes where capillary forces are important than in very large pipes, where very large continuous slugs are
less probable. On the other hand, slug formation in small scale does not necessarily lead to any problems (it
can sometimes be even desired), but slugs in large industrial pipes may lead to failure in piping constructions
due to pressure shocks and heavy vibrations.

23
Pressure drop for two-phase flows is often calculated with such a procedure, where first flow regime is
identified based on superficial velocities of the two phases, then pressure drop is calculated for the two
phases separately (as if they would be flowing without the other phase), and then empirical correlations is
used to combine these information for a two-phase pressure drop and holdup. One classical approach for
this is by Lockhart and Martinelli (not reported here in detail). Many others exist as well. Many of them are
also implemented in process flowsheet simulators, such as Aspen Plus. Naturally care must be taken when
using these correlations, and combined with common sense.

In some cases, the two phases may be separated and fed to separate gas and liquid pipes. This requires a
separate settling tank with liquid level control, but for example when distillation column feed is partially
vaporized liquid, this arrangement could in some cases give more stable feed distributor system inside the
column as splashes due to slugs are avoided.

6.2 Interphase forces


Forces between dispersed phase entities (bubbles, droplets or particles) are needed in order to be able to
predict their movement with respect to the continuous phase. In many cases, the final outcome is formulated
in terms of terminal velocity, i.e. velocity of the particle if it would rise or drop freely in a large volume of
continuous fluid. Even in these cases, the result can be simply a steady state solution of momentum balance
written for an individual particle. The momentum balance relevant for this situation can be written as

d ∋mU (
< V∋θ p , θf (g , FD (32.)
dt

The term on the left is rate of change of momentum, which in practice means acceleration. The first term on
the right hand side is buoyancy, i.e. force resulting from the density differences. This is basically the driving
force for separation: if the densities of the two phases are similar, they flow together. This may be problem
e.g. in extraction where the densities of the two phases may be close to each other. In extraction, it is
important to select the solvent so that there is a density difference between the phases.

The last term on the right is interfacial force. This is actually a combination of several force terms, but in
many cases drag is the most relevant term. Drag can be calculated with definition of the drag coefficient:

1
FD < C D A p θU 2 (33.)
2

Here CD is the drag coefficient, similar in nature than friction coefficients discussed earlier. Ap is the cross-
sectional area of the particle as seen by the continuous phase. For example for spherical bubbles or droplets
of diameter d, cross-sectional area would be Ap = ο/4d2. U is the velocity difference between the particle and
the continuous phase.

There are several empirical correlations and graphs available for the drag coefficient, for example:

24
For this, the following three regime correlation gives reasonable approximation for drag coefficients for
spherical particles with relevant Reynolds numbers in chemical engineering applications:

Re < 1 CD = 24/Re

1 < Re < 103 CD = 18*Re-0.6

103 < Re CD = 0.44

For a particle Reynolds number, the characteristic diameter is always the particle diameter.

When the first of these is inserted into the momentum equation, spherical particle and steady state is
assumed, the so called Stokes law for terminal velocity is obtained. Stokes' law is valid for very small bubbles
or droplets. When the last of the regimes is assumed (constant drag), the so called Newtonian settling law is
obtained. That is valid for relatively large particles, where flow around the particle is turbulent, but boundary
layer is not yet separated from the particle surface (drop at high Re values in the graph). There are several
other correlations also available.

One thing worth noting here is that the previous analysis is valid only for spherical particles, although even
solid particles are often of different shape. For relatively large bubbles and droplets, their shape starts to
deviate from a spherical one. There are also flow regime maps for fluid particle shapes. One such is presented
in the following figure, where bubble shape as a function of its physical properties is given. Based on the
graph, bubble Reynolds number and thus rising velocity can be estimated when its Eotvos and Morton
numbers are known.

25
where

ΧθgD 2
Eo < (34.)
ρ

Χθgλ 4
Mo < 2 3 (35.)
θf ρ

With given physical properties (specified chemical system), a simpler graph relating bubble diameter to its
terminal velocity can be obtained. One such is in the following figure for air-water system. One important
point is that bubble rise velocity depends on whether the liquid is contaminated (contains surfactants) or
not. Fluid inside pure droplets and bubbles circulate with moving interface. This reduces drag (increases
terminal velocity) as compared to solid particles. Even small amounts of surfactants concentrate on gas-liquid
surfaces, and stop surface movement. Besides bubble rise velocity, this has an impact on gas-liquid mass
transfer and bubble coalescence tendency. In practice water contains almost always some surfactants so in
most practical cases it can be considered as contaminated from this perspective (although it would be
perfectly fine as a drinking water).

26
One noteworthy point in gas bubble rise is that due to onset of oscillations, terminal velocity is often rather
constant over a wide range of bubble sizes relevant for chemical engineering applications (from 1 to 20 mm).
For rough calculations, bubbles can often be assumed to rise approximately 20-30 cm/s. The same is not true
for liquid droplets, where terminal velocities for relevant droplet sizes in chemical engineering applications,
such as in extraction, are not constant.

6.3 Fluidization
Fluidization is an operation, where upwards flowing gas or liquid overcomes gravity for individual particles.
Limiting velocity for this is called minimum fluidization velocity. If the flow is slower, the particles settle on
the bottom of the equipment, typically on top of a grid preventing particles to drop to the gas or liquid feeding
system. Above the minimum fluidization velocity, particles start to float freely in the vessel. Typically the
system is relatively well mixed, which is preferable e.g. in cases where catalytic exothermic reaction occur in
the particles. In fluidized mode, it is also possible to continuously remove solids from the vessel. If fluid
velocity is increased further, particles do not stay in the vessel anymore, and pneumatic conveying results in.

Fluidization is extensively used in oil refineries, where the core of modern refineries is FCC (fluid catalytic
cracking) unit. Less valuable heavy oils are converted into light components more suitable to be used as
gasoline compounds. Catalyst in these operations deactivate due to coke formation very rapidly, so catalyst
particles are allowed to flow continuously through the reactor part into a regenerator where coke is burned
away. After this, regenerated catalyst is returned to the reactor.

Fluidization can be used also for other processes, such as in drying. Typical flow patterns that can be found
in fluidization are shown in the following figure.

27
6.4 Two-phase flow in porous media
In some cases, the flow in porous media contains two liquid phases in addition to the solid matrix through
which the fluids flow. For example in trickle bed reactor, gas and liquid flow co-currently through fixed bed
of solid catalyst particles. In these cases, both pressure drop and liquid hold-up in the reactor are relevant
design parameters. In order to predict these, momentum balances for both gas and liquid needs to be
formulated. After simplification, pressure drop correlations for both phases are obtained, with appropriate
interaction terms between gas and liquid, liquid and solid, and perhaps between gas and solid in cases where
part of the solid is not wetted. The calculated pressure drop for both phases needs to be the same in a given
bed. This can be achieved since the volume fraction of the flowing fluids give one additional degree of
freedom. One set of equations suitable for predicting two-phase flow in packed beds is the following.

Pressure drops for both of the phases:

Χp , f GL
< ∗ θG g (36.)
L 

Χp f GL , f LS
< ∗θ g (37.)
L ∋1 ,  ( L

gas-liquid and liquid-solid interactions (fully wetted solid phase assumed here):

f GL < ∋A GL λ G jr ∗ BGL θG jr jr ( (38.)

f LS < ∋1 ,  (∋A LSλ L jL ∗ BLSθ L jL jL ( (39.)

28
Relative velocity of the phases in the pores:

jG 
jr < , jL (40.)
 ∋1 ,  (

Interaction terms:

A GL < E λGL
∋1 , δ (2 (41.)
3δ3d 2p

BGL < E θGL


∋1 , δ ( (42.)
 3δ 3d p

A LS < E λLS
∋1 , δ (2 (43.)
∋1 ,  (3 δ3d 2p

BLS < E θLS


∋1 , δ( (44.)
∋1 ,  (3 δ3d p

This system of equation can be solved iteratively to give pressure drop and void fraction (gas saturation).

6.5 Capillary pressure and surface wetting


One further thing to consider in multi-phase flows is capillary pressure. It affects wetting properties of porous
material, formation of small bubbles (bubble nucleates), and in many other relevant phenomena in chemical
engineering. The following figure illustrates capillary pressure in simple capillaries immersed into liquid, and
how capillary pressure affects fluid saturation in porous media.

29
With narrow capillaries or small particles with small openings in between them, capillary pressure could be
strong. For larger pipes or flow in between large particles, it is less significant. Capillary pressure can be
estimated, if the curvature of the surface is known from the Young-Laplace equation:


pc < (45.)
R

Here R is the radius of curvature, and ρ is the surface tension.

Capillary pressure is often advantageous, as it helps to properly wet solid surfaces, provided the surface
material is such that it is preferably wetted by the liquid. This is important e.g. in distillation or absorption
with packed columns, in coalescers helping to increase drop size before settling, or in trickle bed reactors,
where proper wetting of catalyst is important for good reactivity and to avoid hot spots in the reactor.

7 Mixing
One particularly important process in chemical engineering where fluid flow aspects are important is mixing.
Mixing may be needed e.g. in chemical reactors to get reactants in one or several phases into good contact
with each other, facilitate mass transfer between the phases by increasing interfacial area (breaking large
bubbles or droplets) and preventing settling, or homogenizing products. Although mixing may appear as a
simple operation, it is important to optimize mixing processes in order to minimize energy input.

7.1 Mixing in stirred tanks


In many cases, it is convenient to carry out mixing in vessels with impeller(s). These impellers mainly act to
move (blend) fluid and disperse phases for a good contact. For various purposes, there are a large number
of different impeller and mixing tank geometries. For example laminar mixing requires typically completely
different impeller geometry than turbulent mixing. Mixing of various non-Newtonian fluids may also require
special designs.

Mixing can be characterized with various parameters. The first to be calculated is the Reynolds number. In
stirred tanks, it is calculated based on impeller diameter and its tip speed. Typically the constant factor ο is
not included in the impeller Reynolds number definition, so it becomes

θND 2
Re < (46.)
λ

If impeller Re is above 104, the flow can be considered turbulent, if Re is much below 1000, flow is laminar.
In between there is a transitional regime similarly as in the pipe flow. Also for impellers, these limits cannot
be defined very precisely.

Besides the Reynolds number, perhaps the most important factor characterizing mixing is power input. It can
be estimated when impeller power number is known:

30
P < N P θN 3D 5 (47.)

Again there are graphs available to estimate impeller power numbers. One such is shown below.

Note that for laminar flow, power number is inversely proportional to the impeller Reynolds number, and for
fully turbulent flow, it is practically constant (but different for each impeller type). For example for a
commonly used Rushton turbine (impeller 1), the power number is approximately 5 in the turbulent regime.
Similar behavior was earlier found in drag coefficient in pipes, or drag of particles in free rise or fall in
continuous fluid.

Other important parameters for mixing in stirred tanks are flow from the impeller (based on the pumping
number) and blending time, i.e. time required to achieve certain degree of homogeneity after a tracer feed.
These can be estimated with the following equations:

Q < N Q ND3 (48.)

T1.5 H 0.5
Nπ < 5.2 2 1 / 3 (49.)
D Np

The first of these is a definition of a pumping number N Q, and the second is an empirical correlation to
measured blending times in various stirred tanks. Pumping numbers can be found for different impeller types
in similar graphs and correlations as the power number.

31
7.2 Multiphase mixing in stirred tanks
As discussed earlier, in many chemical engineering applications processes contain multiple phases. This is
also true for mixing in stirred tanks. Some examples are aerated fermenters or other gas-liquid reactors,
stirred extraction columns with or without chemical reaction, crystallization, most polymer reactors, and so
on. In some cases there may be even more than two phases, such as in flotation. In these cases it is important
that the phases are in good contact, does not separate unintentionally, and that the whole vessel contain all
the phases.

For mixing power, the simplest approach is to calculate just effective dispersion viscosity and density, and
use that in one-phase correlations. Especially for gas-liquid mixing, gassing reduces power input by the
impeller, which can be taken into account with specific gassed system power correlations.

There are several empirical correlations available for estimating whether the conditions are suitable for two-
phase mixing. For example for solid-liquid mixing, solid phase should be well suspended. If a significant
fraction of solids lay at the bottom of the vessel, mass transfer is reduced, and in worst cases the particles
may form a single solid bulk, e.g. in crystallizers where the particles may form agglomerates.

The figure below describes the situation where (a) particles are not properly suspended, (b) above so called
just suspended speed, and (c) fully suspended dispersion.

The most important design criterion is to be above just suspended speed, i.e. avoid regime (a). For this, there
are several correlations. One such is by Zwietering:

0.2
D
0.1
Reimp Fr 0.45   X 0.13 < S (50.)
 d p 

where X is the mass ratio of suspended solids * 100 (i.e. %), S is a tank geometry specific parameter, typically
between 3-8 for most typical cases. Typically when these correlations are formulated, much more data is
available from relatively standard geometries. For more exotic designs, these correlations are less reliable.

Another typical multi-phase design problem is gas feed and impeller speed to avoid flooding in aerated
reactors. Impeller flooding is illustrated qualitatively in the next figure:

32
When too much gas is fed to the system, it starts to accumulate behind impeller blades. In that situation
bubbles coalesce into a single gas cavity behind the impeller, and leave that regime as large bubbles with
small mass transfer area. Bubble phase is also poorly distributed in the vessel, as gas typically rises just close
to the impeller axis.

In order to avoid this, there are empirical correlations again available. One such is presented here to predict
onset of flooding:

3 .5
Q ∑D⌡
Fl G < = 30 Fr   (51.)
ND 3 T

It can be seen that flooding occurs if gassing rate is very high compared to the impeller speed.

Another gassing related problem for very large vessels is that gas superficial velocity (volume flow divided by
vessel cross-sectional area) increases in scale-up if gas feed to vessel volume is kept constant. This leads
ultimately to change in flow regime, so that very large vessels may be closer to bubble columns than agitated
vessels. In very large scales, this may lead to different optimal structure than in small scales.

Impeller selection can also partially help to avoid flooding. Instead of flat blades, such as in Rushton turbine,
curved impellers can be used so that cavity behind impeller does not form so easily.

7.3 Static mixers


In some cases, mixing is easier to accomplish in a pipe than in separate vessel. Normal pump is used for power
input, and a static mixer could be inserted in the pipe to enhance mixing. For turbulent flow, the main reason
for static mixers is to reduce time needed for mixing (flows would mix by themselves anyway), but for laminar
flows static mixers may be needed since flow would mix only very slowly with molecular diffusion mechanism.
One (out of many) possible static mixer type is shown in the following figure.

33
8 Computational fluid dynamics (CFD)
Since the Navier-Stokes equations can be solved analytically only in very simple special cases, a numerical
solution is required. One complication in fluid flow modeling compared to many other modeling efforts in
chemical engineering is that the geometry where the interesting flow occurs is often relatively complex
where strong symmetries cannot be assumed. This is very different compared to e.g. reactor or mass transfer
operation modeling, where often zero- or one-dimensional models (e.g. plug flow reactor) can be used with
success. Sometimes radial or other symmetries can be used for fluid flow, or only part of the unit needs to
be modeled (half or one quarter, but 3D). If these are possible, it is always advisable to take advantage of
such symmetries. Note however that although the equipment itself would seem symmetrical, the flow itself
could be unstable in such a way that the symmetry is broken. One example is bubble column, where the
column itself could in some cases be assumed approximately symmetrical, but the bubble swarm swirls
around the column so that the final flow is not only three-dimensional but also time dependent.

In order to solve the fluid flow, the modeled volume is divided into a computational grid. In the following
figure, there are some discrete control volumes to be used in computational fluid dynamics

34
The whole domain must be discretized with elements such as those in this figure. The ones on the left are for
2D modeling, and the others for 3D.

Once the domain is discretized, a numerical scheme is needed to solve the N-S equations along with other
possible equations, such as turbulence model, energy balance or chemical component balances. There are
various methods for this, but perhaps the most often used in the field of CFD is control volume method. It is
also relatively easy to understand, as separate balances are formulated for each element. Fluxes at the
boundaries for all the necessary variables (mass, momentum etc.) are calculated from the cell values with
suitable interpolation schemes. For fluid flow, often so called upwind schemes are used, where information
at the boundaries are taken from the upwind side, i.e. where the flow is coming from. Detailed algorithms
for solving the equations are outside the scope of this course.

There are also other methods that can be used for solving the discretized balances, for example finite element
method (FEM) used by COMSOL software. Origins of FEM are in other fields of physics, and it is used
extensively e.g. in structural mechanics.

8.1 Evaluation of the grid


One of the most crucial aspects in CFD is to evaluate whether the discretization, i.e. the grid, is good for the
purpose. A poor grid may lead to completely erroneous results, and a "too good" grid with excessive number
of grid points leads to very slow solution of the model. Nowadays 3D CFD grids typically contain around 10 4-
106 cells in engineering applications, so that the number of variables is typically at least six times larger than
that (see discussion related to N-S equations and turbulence modeling). Each variable, e.g. new chemical
compound which concentration profiles we wish to follow, increases the total number of variables by the
number of cells.

Even if the number of grid points is optimal, the solution may be slow if the grid topology is poor. This may
be simply a result of a wish to model very small details in large equipment, requiring very small cells near the
small details, and large cells elsewhere. Then the small cells may be crucial to the computational speed
although the total number of cells is reasonable.

In the following figure one typical CFD grid is shown. The modeled piece of equipment is part of a laminar
mixer, where two flows coming from the left are divided and recombined in order to form a lamella kind of
structure for the flow, where diffusion lengths would be short implying fast mixing. These kinds of structures
are used in microprocess technology, where the small scales result in laminar flow. In laminar flows, mixing
is always much slower compared to similar situation with turbulence.

35
In the figure, the edges of the computational cells at the walls are shown. The cells continue and fill the space
where fluid flows. It can be seen that the cell size is not the same everywhere; near the corners there are
smaller cells than elsewhere. This is necessarily due to numerical reasons; otherwise the whole solution
would suffer from inaccuracy. In general, the grid shown in the figure above is not particularly dense, and
still suffers probably from numerical diffusion (one form of inaccuracy smoothing sharp gradients).

One good way of evaluating grid is to check the solutions with different number of cells. Then a suitable
value, such as overall flow rate or total energy dissipated to the system is calculated, and checked whether
this value changes as a function of grid size. In the following figure a typical situation is shown, where the
number of cells is on the x-axis and a modeled property on y-axis. Sometimes there are different ways to
calculate the modeled property, and predictions with the different methods do not give the same result until
the grid is good enough. One example is mixing power in stirred tanks. It can be predicted from the turbulent
energy dissipation averaged over all the cells, or from angular moment imposed to the impeller blade.
Typically the former is converged to the final value with a higher number of cells than the latter. Then
depending on the primary reason for the CFD analysis, the optimal number of cells may also vary.

36
One further issue that needs to be taken into account is the stability of time dependent CFD solution. If the
time step is too long compared to the flow, the solution is not stable. In practice, the flow should not proceed
more than one cell interval at a given time step (the flow should not "jump" over a cell). This limitation is
known as Courant-Friedrichs-Lewy criterion.

8.2 Modeling of equipment with moving parts


In many cases, there are moving parts (walls from the CFD point of view) in the equipment. For example in
stirred tanks impeller moves with respect to the equipment. Then the grid needs to be somehow adjusted to
take this into account. There are two commonly used methods: Multiple Reference Frames (MRF) and moving
grid. In the first, the vessel is divided into two sub-regions, one describing the volume close to the impeller
and one the volume outside it. The impeller region moves with the impeller so that the grid is stagnant with
respect to the impeller. The outside region is stagnant with respect to the other vessel. As the two grids move
with respect to each other, the MRF method needs to distribute information between the two regions.

The moving grid method is, according to its name, such that the grid is adjusted at each time step to the new
geometry. This is very versatile method and can be used for any changing geometry. However, it is
computationally more challenging than MRF. The ideas of the two methods are illustrated in the following
figure:

MRF with two regions Moving grid

8.3 Computational modeling of multiphase systems


Multiphase fluid flow is very typical in chemical engineering applications, which adds complexity to many
other fluid flow modeling tasks for example in mechanical engineering problems. This is the case in
multiphase reactors, separation processes, agglomeration, filtration and so on. In many cases the structure
of the two phase system is not well defined, for example in randomly packed beds (catalyst particles or
random packings in mass transfer operations). In those cases, the region can be assumed porous, and friction
terms for CFD can be taken from the similar terms discussed in the mechanical energy balances.

In other cases, there is a flowing dispersed phase, for example bubbles, droplets or solid particles. In those
cases, there are typically two options. The first is when both cases are modeled with their own momentum
equations, and the interaction between them with drag laws. This is called Euler-Euler, or interpenetrating

37
continua approach. Here the precise locations of the dispersed phase elements (bubbles, droplets...) are not
resolved, but just a volume fraction of each phase is assumed for each computational cell. Another option is
so called Lagrangian approach, or Euler-Lagrange as continuous phase is modeled normally. In this case,
individual particles (or particle swarms) are modeled so that their locations and velocities are known.

Typically Euler-Euler approach is used for high volume fractions and large systems, and Euler-Lagrange for
relatively dilute systems and when particle impacts to solid structures are important. Even for Euler-Euler
approach, approximate fluid element paths can be estimated with tracer particles after a solved CFD
simulation.

Essential in a successful two-phase simulation is correct modeling of interphase momentum transfer.


Typically drag force is the most important, but there are also other possible relevant forces, such as lift force
or virtual mass. Even for the drag force, there are several possible models available, and it is not always self-
evident which one to select. Sensitivity and case studies are often needed if the results cannot be validated
directly based on experimental data.

8.4 Modeling of moving fluid interfaces


In some cases, the location of the fluid-fluid interface is important, so that the interface itself needs to be
resolved. In academic research, bubble or droplet behavior (shape, breakage etc.) can be simulated directly
to get more insight to the system. This is often more of academic interest and to develop more practical
correlations for the modeled phenomenon. Sometimes the interface is also of practical interest, for example
in settling units and in case of slugs or other large fluid elements compared to the equipment size. Also
surface shape may be of interest, e.g. in mixing whether there is such a surface vortex that enters the impeller
regime leading to excessive entrainment of gas to the liquid.

There are different methods to model the moving fluid interface:

- Volume of Fluid (VOF). A volume fraction ("color function") is used to describe fractions of both
phases in a cell. The interface lies in cells where this variable is not 0 or 1. Various algorithms exist
for trying to avoid interface smearing (it should be sharp).
- Level-set. A smooth scalar function is added to describe the interface location. The scalar is moved
according to a transport equation.
- Front-tracking. The phase interface is tracked with a set of marker particles.

In all these cases the grid should be dense enough to capture the interface curvature so that surface tension
related forces can be taken into account properly.

38
9 Flow measurement
In order to control processes, various measurements are needed. The most common ones are perhaps
temperature and pressure measurements, but there are several equally important, such as fluid level
measurement and flow rate measurements.

There are numerous flow measurement devices based on different operating principles. One possible
classification is the following (in parentheses some flowmeter types in each category):

1. Flowmeters with wetted moving parts (positive displacement, hydraulic Wheatstone bridge, turbine,
variable area)

- sensitive to wear, mechanical failure can occur


- typically only for clean fluids

2. Flowmeters with no wetted moving parts (differential pressure, oscillatory, target, thermal)

- plugging or wear can occur


- adds pressure drop

3. Obstructionless flowmeters (Coriolis mass, magnetic, ultrasonic)

- basically a subset of 2, but fluid flows through freely

4. Flowmeters with sensors mounted external to the pipe (clamp-on ultrasonic, correlation)

- Sensor material not sensitive to fluid properties

39
10 Symbols
A area (m2)
c concentration (mol/m3)
CD drag coefficient ()
Cp, Cv specific heat capacities at constant pressure and volume (J/molK) or (J/kgK)
D diameter (m)
DAB diffusion coefficient (m2/s)
dp particle size (m)
F force (N)
f force term (various) (N)
FD interfacial force (N)
g gravitational acceleration constant 9.81 (m/s2)
H stirred tank height (m)
hf friction loss (J/kg)
j superficial velocity (m/s)
k ratio of specific heat capacities ()
k turbulent kinetic energy (J/kg = m2/s2)
k permeability (m2)
L length (m)
m, n parameters of viscosity models ()
n polytropic exponent ()
N impeller speed (1/s)
NP power number ()
NQ flow number ()
p pressure (Pa)
P power W
Q flow rate (m3/s)
R gas constant 8.314 (J/molK)
r radius (m)
S source term (various)
S tank geometry parameter ()
t time (s)
T temperature (K)
T stirred tank diameter (m)
U velocity (m/s)
v velocity (m/s)
V volume (m3)
W pumping work (J/kg)
Wcompr compressor work (J/mol)
x length coordinate (m)
X mass ratio of suspended solids (%)
y length coordinate (m)
z height position, coordinate (m)

40
 flow profile correction factor ()
 gas saturation (gas fraction in void space) ()
ρ surface (interface) tension (N/m)
θ density (kg/m3)
ω pipe friction factor (Darcy) ()
κ thermal conductivity (W/mK)
δ porosity ()
σ shear stress (N/m2 = Pa)
σγ Kolmogorov time scale (s)
γ Kolmogorov length scale (m)
γ efficiency ()
ψ local friction coefficient ()
µ viscosity (kg/ms = Pas)

41

You might also like