You are on page 1of 53

“MODELLING & SIMULATION OF BINARY

DISTILLATION COLUMN”

CONTE
NTS

• Introduction
• Vapour Liquid Equilibrium
• Types of Distillation
1. Batch Distillation
2. Continuous Distillation
• Simple Distillation
• Flash Evaporation
• Fractional Distillation
• Types of Azeotropes
• Separation of Azeotropes
• Steam Distillation
• Vacuum Distillation
• Extractive Distillation
• Theoretical plates
• Methods of calculating no. of stages
1. Fenske Equation
2. McCabe-Thiele Method
• Modelling of McCabe-Thiele Method
• Assumptions of McCabe-Thiele Method
• Introduction to Simulation
• Flow Chart of Simulation Program
• Problem
• Discussion & Conclusion
• Appendix
• Bibliography

Introduction
A process in which a liquid or vapour mixture of two or more substances is
separated into its component fractions of desired purity, by the application
and removal of heat.

Or in other words: Distillation is a widely used method for separating


mixtures based on differences in the conditions required to change the phase
of components of the mixture. To separate a mixture of liquids, the liquid
can be heated to force components, which have different boiling points, into
the gas phase. The gas is then condensed back into liquid form and collected.
Repeating the process on the collected liquid to improve the purity of the
product is called double distillation. Although the term is most commonly
applied to liquids, the reverse process can be used to separate gases by
liquefying components using changes in temperature and/or pressure.
Distillation is used for many commercial processes, such as production of
gasoline, distilled water, xylene, alcohol, paraffin, kerosene, and many other
liquids. Types of distillation include simple distillation (described here),
fractional distillation (different volatile 'fractions' are collected as they are
produced), and destructive distillation (usually, a material is heated so that it
decomposes into compounds for collection).

Distillation is based on the fact that the vapour of a boiling mixture will be
richer in the components that have lower boiling points.
Therefore, when this vapour is cooled and condensed, the condensate will
contain more volatile components. At the same time, the original mixture
will contain more of the less volatile material. Distillation columns are
designed to achieve this separation efficiently.
Although many people have a fair idea what “distillation” means, the
important aspects that seem to be missed from the manufacturing point of
view are that:
• Distillation is the most common separation technique.
• It consumes enormous amounts of energy, both in terms of cooling
and heating requirements.
• It can contribute to more than 50% of plant operating costs.

The best way to reduce operating costs of existing units, is to improve their
efficiency and operation via process optimization and control. To achieve
this improvement, a thorough understanding of distillation principles and
how distillation systems are designed is essential.

The Distillation of the Crude Oil in Oil Refinery can be represented by:
Vapor-Liquid Equilibrium
Vapor-liquid equilibrium, abbreviated as VLE by some, is a condition where
a liquid and its vapor (gas phase) are in equilibrium with each other, a
condition or state where the rate of evaporation (liquid changing to vapor)
equals the rate of condensation (vapor changing to liquid) on a molecular
level such that there is no net (overall) vapor-liquid inter conversion.
Although in theory equilibrium takes forever to reach, such an equilibrium is
practically reached in a relatively closed location if a liquid and its vapor are
allowed to stand in contact with each other long enough with no interference
or only gradual interference from the outside.

VLE Data Introduction


The concentration of a vapor in contact with its liquid, especially at
equilibrium, is often given in terms of vapor pressure, which could be
a partial pressure (part of the total gas pressure) if any other gas(es) are
present with the vapor. The equilibrium vapor pressure of a liquid is usually
very dependent on temperature. At vapor-liquid equilibrium, a liquid with
individual components (compounds) in certain concentrations will have an
equilibrium vapor in which the concentrations or partial pressures of the
vapor components will have certain set values depending on all of the liquid
component concentrations and the temperature. This fact is true in reverse
also; if a vapor with components at certain concentrations or partial
pressures is in vapor-liquid equilibrium with its liquid, then the component
concentrations in the liquid will be set dependent on the vapor
concentrations, again also depending on the temperature. The equilibrium
concentration of each component in the liquid phase is often different from
its concentration (or vapor pressure) in the vapor phase, but there is a
correlation. Such VLE concentration data is often known or can be
determined experimentally for vapor-liquid mixtures with various
components. In certain cases such VLE data can be determined or
approximated with the help of certain theories such as Raoult's
Law, Dalton's Law, and/or Henry's Law.
Such VLE information is useful in designing columns for distillation,
especially fractional distillation, which is a particular specialty of chemical
engineers. Distillation is a process used to separate or partially separate
components in a mixture by boiling (vaporization) followed
by condensation. Distillation takes advantage of differences in
concentrations of components in the liquid and vapor phases.
In mixtures containing two or more components where their concentrations
are compared in the vapor and liquid phases, concentrations of each
component are often expressed as mole fractions. A mole fraction is number
of moles of a given component in an amount of mixture in a phase (either
vapor or liquid phase) divided by the total number of moles of all
components in that amount of mixture in that phase.
Binary mixtures are those having two components. Three-component
mixtures could be called ternary mixtures. There can be VLE data for
mixtures with even more components, but such data becomes copious and is
often hard to show graphically. VLE data is often shown at a certain overall
pressure, such as 1 atm or whatever pressure a process of interest is
conducted at. When at a certain temperature, the total of partial pressures of
all the components becomes equal to the overall pressure of the system such
that vapors generated from the liquid displace any air or other gas which
maintained the overall pressure, the mixture is said to boil and the
corresponding temperature is the boiling point (This assumes excess
pressure is relieved by letting out gases to maintain a desired total pressure).
A boiling point at an overall pressure of 1 atm is called the normal boiling
point.

Thermodynamic Description of Vapor-Liquid Equilibrium


The field of thermodynamics describes when vapor-liquid equilibrium is
possible, and its properties. Much of the analysis depends on whether the
vapor and liquid consist of a single component, or if they are mixtures.

Pure (single-component) systems

If the liquid and vapor are pure, in that they consist of only one molecular
component and no impurities, then the equilibrium state between the two
phases is described by the following equations:

And
where and are the pressures within the liquid and vapor,
and are the temperatures within the liquid and vapor, and
and are the molar Gibbs free energies (units of energy per amount of
substance) within the liquid and vapor, respectively.[4] In other words, the
temperature, pressure and molar Gibbs free energy are the same between the
two phases when they are at equilibrium.
An equivalent, more common way to express the vapor-liquid equilibrium
condition in a pure system is by using the concept of fugacity. Under this
view, equilibrium is described by the following equation:

where and are the fugacities of the liquid and


vapor, respectively, at the system temperature and pressure .[5] Using
fugacity is often more convenient for calculation, given that the fugacity of
the liquid is, to a good approximation, pressure-independent,[6] and it is
often convenient to use the quantity , the dimensionless fugacity
coefficient, which is 1 for an ideal gas.

Multicomponent systems

In a multicomponent system, where the vapor and liquid consist of more


than one type of molecule, describing the equilibrium state is more
complicated. For all components in the system, the equilibrium state
between the two phases is described by the following equations:

where and are the temperature and pressure for each phase, and
and are the partial molar Gibbs free energy also called chemical
potential (units of energy per amount of substance) within the liquid and
vapor, respectively, for each phase. The partial molar Gibbs free energy is
defined by:
where is the (extensive) Gibbs free energy, and is the amount of
substance of component .

Types Of Distillation Columns


There are many types of distillation columns, each designed to perform
specific types of separations, and each design differs in terms of complexity.
One way of classifying distillation column type is to look at how they are
operated. Thus we have:
1. Batch and
2. Continuous columns.

Batch Columns
In batch operation, the feed to the column is introduced batch-wise. That
is, the column is charged with a 'batch' and then the distillation process is
carried out. When the desired task is achieved, a next batch of feed is
introduced.

Heating an ideal mixture of two volatile substances A and B (with A having


the higher volatility, or lower boiling point) in a batch distillation setup (such
as in an apparatus depicted in the opening figure) until the mixture is boiling
results in a vapor above the liquid which contains a mixture of A and B. The
ratio between A and B in the vapor will be different from the ratio in the
liquid: the ratio in the liquid will be determined by how the original mixture
was prepared, while the ratio in the vapor will be enriched in the more
volatile compound, A (due to Raoult's Law, see above). The vapor goes
through the condenser and is removed from the system. This in turn means
that the ratio of compounds in the remaining liquid is now different from the
initial ratio (i.e. more enriched in B than the starting liquid).
The result is that the ratio in the liquid mixture is changing, becoming richer
in component B. This causes the boiling point of the mixture to rise, which
in turn results in a rise in the temperature in the vapor, which results in a
changing ratio of A : B in the gas phase (as distillation continues, there is an
increasing proportion of B in the gas phase). This results in a slowly
changing ratio A : B in the distillate.
If the difference in vapor pressure between the two components A and B is
large (generally expressed as the difference in boiling points), the mixture in
the beginning of the distillation is highly enriched in component A, and
when component A has distilled off, the boiling liquid is enriched in
component B.

Continuous Columns
In contrast, continuous columns process a continuous feed stream. No
interruptions occur unless there is a problem with the column or surrounding
process units. They are capable of handling high throughputs and are the
most common of the two types. We shall concentrate only on this class of
columns.

Continuous distillation is an ongoing distillation in which a liquid mixture is


continuously (without interruption) fed into the process and separated
fractions are removed continuously as output streams as time passes during
the operation. Continuous distillation produces at least two output fractions,
including at least one volatile distillate fraction, which has boiled and been
separately captured as a vapor condensed to a liquid. There is always a
bottoms (or residue) fraction, which is the least volatile residue that has not
been separately captured as a condensed vapor.
Continuous distillation differs from batch distillation in the respect that
concentrations should not change over time. Continuous distillation can be
run at a steady state for an arbitrary amount of time. Given a feed of in a
specified composition, the main variables that affect the purity of products in
continuous distillation are the reflux ratio and the number of theoretical
equilibrium stages (practically, the number of trays or the height of packing).
Reflux is a flow from the condenser back to the column, which generates a
recycle that allows a better separation with a given number of trays.
Equilibrium stages are ideal steps where compositions achieve vapor-liquid
equilibrium, repeating the separation process and allowing better separation
given a reflux ratio. A column with a high reflux ratio may have fewer
stages, but it refluxes a large amount of liquid, giving a wide column with a
large holdup. Conversely, a column with a low reflux ratio must have a large
number of stages, thus requiring a taller column.
Continuous distillation requires building and configuring dedicated
equipment. The resulting high investment cost restricts its use to the large
scale.

Types of Continuous Columns


Continuous columns can be further classified according to:

1. The nature of the feed that they are processing

• Binary column - feed contains only two components.


• Multi-component column - feed contains more than two
components.

2. The number of product streams they have

• Multi-product column - column has more than two product


streams.

3. Where the extra feed exits when it is used to help with the
separation

• Extractive distillation - where the extra feed appears in the


bottom product stream.
• Azeotropic distillation - where the extra feed appears at the top
product stream.

4. The type of column internals

• Tray column - where trays of various designs are used to hold


up the liquid to provide better contact between vapour and
liquid, hence better separation.
• Packed column - where instead of trays, 'Packing' are used to
enhance contact between vapour and liquid.
Simple Distillation

In simple distillation, all the hot vapors produced are immediately channeled
into a condenser which cools and condenses the vapors. Therefore, the
distillate will not be pure its composition will be identical to the composition
of the vapors at the given temperature and pressure, and can be computed
from Raoult's law.
As a result, simple distillation is usually used only to separate liquids whose
boiling points differ greatly (rule of thumb is 25 °C) [25] or to separate
liquids from in volatile solids or oils. For these cases, the vapor pressures of
the components are usually sufficiently different that Raoult's law may be
neglected due to the insignificant contribution of the less volatile
component. In this case, the distillate may be sufficiently pure for its
intended purpose.

The liquid mixture that is to be processed is known as the feed and this is
introduced usually somewhere near the middle of the column to a tray
known as the feed tray. The feed tray divides the column into a top
(enriching or rectification) section and a bottom (stripping) section. The feed
flows down the column where it is collected at the bottom in the reboiler.
Heat is supplied to the reboiler to generate vapour. The source of heat input
can be any suitable fluid, although in most chemical plants this is normally
steam. In refineries, the heating source may be the output streams of other
columns. The vapour raised in the reboiler is re-introduced into the unit at
the bottom of the column. The liquid removed from the reboiler is known as
the bottoms product or simply, bottoms.

The vapour moves up the column, and as it exits the top of the unit, it is
cooled by a condenser. The condensed liquid is stored in a holding vessel
known as the reflux drum. Some of this liquid is recycled back to the top of
the column and this is called the reflux. The condensed liquid that is
removed from the system is known as the distillate or top product.
Thus, there are internal flows of vapour and liquid within the column as well
as external flows of feeds and product streams, into and out of the column.

Flash Evaporation
Flash (or partial) evaporation is the partial vaporization that occurs when
a saturated liquid stream undergoes a reduction in pressure by passing
through a throttling valve or other throttling device. This process is one of
the simplest unit operations. If the throttling valve or device is located at the
entry into a pressure vessel so that the flash evaporation occurs within the
vessel, then the vessel is often referred to as a flash drum.

If the saturated liquid is a single-component liquid (for example,


liquid propane or liquid ammonia), a part of the liquid immediately "flashes"
into vapor. Both the vapor and the residual liquid are cooled to the saturation
temperature of the liquid at the reduced pressure. This is often referred to as
"auto-refrigeration" and is the basis of most conventional vapor compression
refrigeration systems.
If the saturated liquid is a multi-component liquid (for example, a mixture
of propane, isobutane and normal butane), the flashed vapor is richer in the
more volatile components than is the remaining liquid.

Flash Evaporation of a Single-Component Liquid


The flash evaporation of a single-component liquid is an isentropic i.e.,
constant entropy) process and is often referred to as an adiabatic flash. The
following equation, derived from a simple heat balance around the throttling
valve or device, is used to predict how much of a single-component liquid is
vaporized.

X = 100 (HuL – HdL) ÷ (HdV – HdL)


where:
X = weight percent vaporized

HuL = upstream liquid enthalpy at upstream temperature and


pressure, J/kg
HdV = flashed vapor enthalpy at downstream pressure and
corresponding saturation temperature, J/kg
HdL = residual liquid enthalpy at downstream pressure and
corresponding saturation temperature, J/kg

If the enthalpy data required for the above equation is unavailable,


then the following equation may be used.

X = 100 · cp (Tu – Td) ÷ Hv

where:
X = weight percent vaporized

cp = liquid specific heat at upstream temperature and pressure,


J/(kg °C)
Tu = upstream liquid temperature, °C

Td = liquid saturation temperature corresponding to the


downstream pressure, °C
Hv = liquid heat of vaporization at downstream pressure and
corresponding saturation temperature, J/kg

(Note: The words "upstream" and "downstream" refer to before and after the
liquid passes through the throttling valve or device.)

This type of flash evaporation is used in the desalination of brackish water or


ocean water by "Multi-Stage Flash Distillation." The water is heated and
then routed into a reduced-pressure flash evaporation "stage" where some of
the water flashes into steam. This steam is subsequently condensed into salt-
free water. The residual salty liquid from that first stage is introduced into a
second flash evaporation stage at a pressure lower than the first stage
pressure. More water is flashed into steam which is also subsequently
condensed into more salt-free water. This sequential use of multiple flash
evaporation stages is continued until the design objectives of the system are
met. A large part of the world's installed desalination capacity uses multi-
stage flash distillation. Typically such plants have 24 or more sequential
stages of flash evaporation.

Equilibrium Flash of a Multi-Component Liquid

The equilibrium flash of a multi-component liquid may be visualized as a


simple distillation process using a single equilibrium stage. It is very
different and more complex than the flash evaporation of single-component
liquid. For a multi-component liquid, calculating the amounts of flashed
vapor and residual liquid in equilibrium with each other at a given
temperature and pressure requires a trial-and-error iterative solution. Such a
calculation is commonly referred to as an equilibrium flash calculation. It
involves solving the Rachford-Rice equation:

Where:
zi is the mole fraction of component i in the feed liquid (assumed to be
known);
β is the fraction of feed that is vaporised;
Ki is the equilibrium constant of component i.
The equilibrium constants Ki are in general functions of many parameters,
though the most important is arguably temperature; they are defined as:

Where:
xi is the mole fraction of component i in liquid phase;
yi is the mole fraction of component i in gas phase.

Once the Rachford-Rice equation has been solved for β, the


compositions xi and yi can be immediately calculated as:
The Rachford-Rice equation can have multiple solutions for β, at most one
of which guarantees that all xi and yi will be positive. In particular, if there is
only one β for which:

Then that β is the solution; if there are multiple such β's, it means that
either Kmax<1 or Kmin>1, indicating respectively that no gas phase can be
sustained (and therefore β=0) or conversely that no liquid phase can exist
(and therefore β=1).

It is possible to use Newton's method for solving the above Rachford-Rice


equation, but there is a risk of converging to the wrong value of β; it is
important to initialize the solver to a sensible initial value, such as
(βmax+βmin)/2 (which is however not sufficient: Newton's method makes no
guarantees on stability), or, alternatively, use a bracketing solver such as the
bisection or the Brent method, which are guaranteed to converge but can be
slower.
The equilibrium flash of multi-component liquids is very widely utilized
in petroleum refineries, petrochemical and chemical plants and natural gas
processing plants.

Fractional Distillation
Fractional distillation is the separation of a mixture into its component parts,
or fractions, such as in separating chemical compounds by their boiling
point by heating them to a temperature at which several fractions of the
compound will evaporate. It is a special type of distillation. Generally the
component parts boil at less than 25 °C from each other under a pressure of
one atmosphere (atm). If the difference in boiling points is greater than
25 °C, a simple distillation is used.

Using the Phase Diagram


If you boil a liquid mixture C1, you will get a vapour with composition C2,
which you can condense to give a liquid of that same composition (the pale
blue lines). If you reboil that liquid C2, it will give a vapour with
composition C3. Again you can condense that to give a liquid of the same
new composition (the red lines).
Reboiling the liquid C3 will give a vapour still richer in the more volatile
component B (the green lines). You can see that if you were to do this once
or twice more, you would be able to collect a liquid which was virtually pure
B. The secret of getting the more volatile component from a mixture of
liquids is obviously to do a succession of boiling-condensing-reboiling
operations. It isn't quite so obvious how you get a sample of pure A out of
this. That will become clearer in a while.

The Vapour
This new vapour will again move further up the fractionating column until it
gets to a temperature where it can condense. Then the whole process repeats
itself.
Each time the vapour condenses to a liquid, this liquid will start to trickle
back down the column where it will be reboiled by up-coming hot vapour.
Each time this happens the new vapour will be richer in the more volatile
component.
The aim is to balance the temperature of the column so that by the time
vapour reaches the top after huge numbers of condensing and reboiling
operations, it consists only of the more volatile component - in this case, B.
Whether or not this is possible depends on the difference between the
boiling points of the two liquids. The closer they are together, the longer the
column has to be.

The Liquid
So what about the liquid left behind at each reboiling? Obviously, if the
vapour is richer in the more volatile component, the liquid left behind must
be getting richer in the other one.
As the condensed liquid trickles down the column constantly being reboiled
by up-coming vapour, each reboiling makes it richer and richer in the less
volatile component - in this case, A. By the time the liquid drips back into
the flask, it will be very rich in A indeed.
So, over time, as B passes out of the top of the column into the condenser,
the liquid in the flask will become richer in A. If you are very, very careful
over temperature control, eventually you will have separated the mixture
into B in the collecting flask and A in the original flask.
Finally, what is the point of the packing in the column?
To make the boiling-condensing-reboiling process as effective as possible, it
has to happen over and over again. By having a lot of surface area inside the
column, you aim to have the maximum possible contact between the liquid
trickling down and the hot vapour rising.
If you didn't have the packing, the liquid would all be on the sides of the
condenser, while most of the vapour would be going up the middle and
never come into contact with it.

Azeotrope
An Azeotrope (pronounced /ay-ZEE-ə-trope/) is a mixture of two or more
liquids (chemicals) in such a ratio that its composition cannot be changed by
simple distillation. This occurs because, when an azeotrope is boiled, the
resulting vapor has the same ratio of constituents as the original mixture.
Because their composition is unchanged by distillation, azeotropes are also
called (especially in older texts) constant boiling mixtures. The
word azeotrope is derived from the Greek words ζέειν (boil) and τρόπος
(change) combined with the prefix α- (no) to give the overall meaning, “no
change on boiling.”
Types of Azeotropes
Each azeotrope has a characteristic boiling point. The boiling point of an
azeotrope is either less than the boiling points of any of its constituents (a
positive azeotrope), or greater than the boiling point of any of its
constituents (a negative azeotrope).
A well known example of a positive azeotrope is 95.6% ethanol and 4.4%
water (by weight). Ethanol boils at 78.4°C, water boils at 100°C, but the
azeotrope boils at 78.1°C, which is lower than either of its constituents.
Indeed 78.1°C is the minimum temperature at which any ethanol/water
solution can boil. In general, a positive azeotrope boils at a lower
temperature than any other ratio of its constituents. Positive azeotropes are
also called minimum boiling mixtures.
An example of a negative azeotrope is hydrochloric acid at a concentration
of 20.2% hydrogen chloride and 79.8% water (by weight). Hydrogen
chloride boils at –84°C and water at 100°C, but the azeotrope boils at 110°C,
which is higher than either of its constituents. The maximum temperature at
which any hydrochloric acid solution can boil is 110°C. In general, a
negative azeotrope boils at a higher temperature than any other ratio of its
constituents. Negative azeotropes are also called maximum boiling mixtures.
Azeotropes consisting of two constituents, such as the two examples above,
are called binary azeotropes. Those consisting of three constituents are
called ternary azeotropes. Azeotropes of more than three constituents are
also known.
More than 18,000 azeotropic mixtures have been documented.
Combinations of solvents that do not form an azeotrope when mixed in any
proportion are said to be zeotropic.
When running a binary distillation it is often helpful to know the azeotropic
composition of the mixture.

Separation of Azeotrope Constituents


Distillation is one of the primary tools that chemists and chemical engineers
use to separate mixtures into their constituents. Because distillation cannot
separate the constituents of an azeotrope, the separation of azeotropic
mixtures (also called azeotrope breaking) is a topic of considerable
interest. Indeed this difficulty led some early investigators to believe that
azeotropes were actually compounds of their constituents. But there are two
reasons for believing that this is not the case. One is that the molar ratio of
the constituents of an azeotrope is not generally the ratio of small integers.
For example, the azeotrope formed by water and acetonitrile contains 2.253
moles of acetonitrile for each mole of water. A more compelling reason for
believing that azeotropes are not compounds is, as discussed in the last
section, that the composition of an azeotrope can be affected by pressure.
Contrast that with a true compound, carbon dioxide for example, which is
two moles of oxygen for each mole of carbon no matter what pressure the
gas is observed at. That azeotropic composition can be affected by pressure
suggests a means by which such a mixture can be separated.

Azeotropic Distillation
Other methods of separation involve introducing an additional agent, called
an Entrainer, that will affect the volatility of one of the azeotrope
constituents more than another. When an entrainer is added to a binary
azeotrope to form a ternary azeotrope, and the resulting mixture distilled, the
method is called azeotropic distillation. The best known example is adding
benzene or cyclohexane to the water/ethanol azeotrope. With cyclohexane as
the entrainer, the ternary azeotrope is 7% water, 17% ethanol, and 76%
cyclohexane, and boils at 62.1°C. Just enough cyclohexane is added to the
water/ethanol azeotrope to engage all of the water into the ternary azeotrope.
When the mixture is then boiled, the azeotrope vaporizes leaving a residue
composed almost entirely of the excess ethanol.

Chemical Action Separation


Another type of entrainer is one that has a strong chemical affinity for one of
the constituents. Using again the example of the water/ethanol azeotrope, the
liquid can be shaken with calcium oxide, which reacts strongly with water to
form the nonvolatile compound, calcium hydroxide. Nearly all of the
calcium hydroxide can be separated by filtration and the filtratere distilled to
obtain nearly pure ethanol.
A more extreme example is the azeotrope of 1.2% water with 98.8% diethyl
ether. Ether holds the last bit of water so tenaciously that only a very
powerful desiccant such as sodium metal added to the liquid phase can result
in completely dry ether.
Anhydrous calcium chloride is used as a desiccant for drying a wide variety
of solvents since it is inexpensive and does not react with most non
aqueous solvents. Chloroform is an example of a solvent that can be
effectively dried using calcium chloride.

Distillation using a Dissolved Salt


When a salt is dissolved in a solvent, it always has the effect of raising the
boiling point of that solvent - that is it decreases the volatility of the solvent.
When the salt is readily soluble in one constituent of a mixture but not in
another, the volatility of the constituent in which it is soluble is decreased
and the other constituent is unaffected. In this way, for example, it is
possible to break the water/ethanol azeotrope by dissolving potassium
acetate in it and distilling the result.

Examples of azeotropes

Proportions are by weight:

• nitric acid (68%) / water, boils at 120.5°C at 1 atm (negative


azeotrope)
• perchloric acid (28.4%) / water, boils at 203°C (negative azeotrope)
• hydrofluoric acid (35.6%) / water, boils at 111.35°C (negative
azeotrope)
• ethanol (96%) / water, boils at 78.1°C
• sulfuric acid (98.3%) / water, boils at 338°C
• acetone / methanol / chloroform form an intermediate boiling (saddle)
azeotrope
• diethyl ether (33%) / halothane (66%) a mixture once commonly used
in anaesthesia.
• benzene / hexafluorobenzene forms a double binary azeotrope.
Complex Azeotrope Systems
The rules for positive and negative azeotropes apply to all the examples
discussed so far. But there are some examples that don't fit into the
categories of positive or negative azeotropes. The best known of these is the
ternary azeotrope formed by 30% acetone, 47%chloroform, and
23% methanol, which boils at 57.5°C. Each pair of these constituents forms
a binary azeotrope, but chloroform/methanol and acetone/methanol both
form positive azeotropes while chloroform/acetone forms a negative
azeotrope. The resulting ternary azeotrope is neither positive nor negative.
Its boiling point falls between the boiling points of acetone and chloroform,
so it is neither a maximum nor a minimum boiling point. This type of system
is called a Saddle Azeotrope. Only systems of three or more constituents
can form saddle azeotropes.

A rare type of complex binary azeotrope is one where the boiling point and
condensation point curves touch at two points in the phase diagram. Such a
system is called a double azeotrope, and will have two azeotropic
compositions and boiling points. An example is water and N-
methylethylenediamine.
Steam Distillation
Steam distillation is a special type of distillation (a separation process)
for temperature sensitive materials like natural aromatic compounds.
Many organic compounds tend to decompose at high sustained temperatures.
Separation by normal distillation would then not be an option, so water
or steam is introduced into the distillation apparatus. By adding water or
steam, the boiling points of the compounds are depressed, allowing them to
evaporate at lower temperatures, preferably below the temperatures at which
the deterioration of the material becomes appreciable. If the substances to be
distilled are very sensitive to heat, steam distillation can also be combined
with vacuum distillation. After distillation the vapors are condensed as usual,
usually yielding a two-phase system of water and the organic compounds,
allowing for simple separation.

Principle
When a mixture of two practically immiscible liquids is heated while being
agitated to expose the surfaces of both the liquids to the vapor phase, each
constituent independently exerts its own vapor pressure as a function of
temperature as if the other constituent were not present. Consequently, the
vapor pressure of the whole system increases. Boiling begins when the sum
of the partial pressures of the two immiscible liquids just exceeds
the atmospheric pressure (approximately 101 kPa at sea level). In this way,
many organic compounds insoluble in water can be purified at a temperature
well below the point at which decomposition occurs. For example, the
boiling point of bromobenzene is 156 °C and the boiling point of water is
100 °C, but a mixture of the two boils at 95 °C. Thus, bromobenzene can be
easily distilled at a temperature 61 C° below its normal boiling point.
Applications
Steam distillation is employed in the manufacture of essential oils, for
instance, perfumes. In this method, steam is passed through the plant
material containing the desired oils. It is also employed in the synthetic
procedures of complex organic compounds. Eucalyptus oil and orange
oil are obtained by this method on the industrial scale.
Steam distillation is also widely used in petroleum
refineries and petrochemical plants where it is commonly referred to as
"steam stripping".
Other industrial uses of steam distillation include the production of
consumer food products such as sprayable or aerosolized condiments such as
sprayable mayonnaise.

Vacuum Distillation
Vacuum distillation is a method of distillation whereby the pressure above
the liquid mixture to be distilled is reduced to less than its vapor
pressure(usually less than atmospheric pressure) causing evaporation of the
most volatile liquid(s) (those with the lowest boiling points). This distillation
method works on the principle that boiling occurs when the vapor pressure
of a liquid exceeds the ambient pressure. Vacuum distillation is used with or
without heating the solution.
Applications
Laboratory-scale vacuum distillation is used when liquids to be distilled
have high atmospheric boiling points or chemically change at temperatures
near their atmospheric boiling points. Temperature sensitive materials (such
as beta carotene) also require vacuum distillation to remove solvents from
the mixture without damaging the product. Another reason vacuum
distillation is used is that compared to steam distillation there is a lower level
of residue build up. This is important in commercial applications
where temperature transfer is produced using heat exchangers.
Vacuum distillation is sometimes referred to as low temperature distillation.
Typical industrial applications utilize the heat pump cycle to maximize
efficiency. This type of distillation is in use in the oil industry where
common ASTM standards are D1160, D2892, D5236. These standards
describe typical applications of vacuum distillation at pressures of about 1-
100 mbar. Pilot plants up to 60 L can be built in accordance with these
standards. Industrial-scale vacuum distillation has several advantages. Close
boiling mixtures may require many equilibrium stages to separate the key
components. One tool to reduce the number of stages needed is to utilize
vacuum distillation.[6] Vacuum distillation columns (as depicted in the
drawing to the right) typically used in oil refineries have diameters ranging
up to about 14 meters (46 feet), heights ranging up to about 50 meters (164
feet), and feed rates ranging up to about 25,400 cubic meters per day
(160,000 barrels per day).
Vacuum distillation increases the relative volatility of the key components in
many applications. The higher the relative volatility, the more separable are
the two components; this connotes fewer stages in a distillation column in
order to effect the same separation between the overhead and bottoms
products. Lower pressures increase relative volatilities in most systems.
A second advantage of vacuum distillation is the reduced temperature
requirement at lower pressures. For many systems, the products degrade or
polymerize at elevated temperatures.
Vacuum distillation can improve a separation by:

• Prevention of product degradation or polymer formation because of


reduced pressure leading to lower tower bottoms temperatures,
• Reduction of product degradation or polymer formation because of
reduced mean residence time especially in columns using
packing rather than trays.
• Increasing capacity, yield, and purity.
Another advantage of vacuum distillation is the reduced capital cost, at the
expense of slightly more operating cost. Utilizing vacuum distillation can
reduce the height and diameter, and thus the capital cost of a distillation
column.

Extractive Distillation
Extractive distillation is similar to azeotropic distillation, except in this case
the entrainer is less volatile than any of the azeotrope's constituents. For
example, the azeotrope of 20%acetone with 80% chloroform can be broken
by adding water and distilling the result. The water forms a separate layer in
which the acetone preferentially dissolves. The result is that the distillate is
richer in chloroform than the original azeotrope.

Theoretical Plate
A theoretical plate in many separation processes is a hypothetical zone or
stage in which two phases, such as the liquid and vapor phases of a
substance, establish an equilibrium with each other. Such equilibrium stages
may also be referred to as an equilibrium stage or a theoretical tray. The
performance of many separation processes depends on having a series of
equilibrium stages and is enhanced by providing more such stages. In other
words, having more theoretical plates increases the efficacy of the separation
process be it either a distillation, absorption, chromatographic, adsorption or
similar process.

Applications
The concept of theoretical plates and trays or equilibrium stages is used in
the design of many different types of separation.

In Distillation Columns
The concept of theoretical plates in designing distillation processes has been
discussed in many reference texts. Any physical device that provides good
contact between the vapor and liquid phases present in industrial-
scale distillation columns or laboratory-scale glassware distillation columns
constitutes a "plate" or "tray". Since an actual, physical plate is rarely a
100% efficient equilibrium stage, the number of actual plates is more than
the required theoretical plates.

where:
Na = the number of actual, physical plates or trays
Nt = the number of theoretical plates or trays
E = the plate or tray efficiency

So-called bubble-cap or valve-cap trays are examples of the vapor and liquid
contact devices used in industrial distillation columns. Another example of
vapor and liquid contact devices are the spikes in laboratory Vigreux
fractionating columns.
The trays or plates used in industrial distillation columns are fabricated of
circular steel plates and usually installed inside the column at intervals of
about 60 to 75 cm (24 to 30 inches) up the height of the column. That
spacing is chosen primarily for ease of installation and ease of access for
future repair or maintenance.
Typical bubble cap trays used in industrial distillation columns
For example, a very simple tray would be a perforated tray. The desired
vapor and liquid contacting would occur as the vapor flowing upwards
through the perforations would contact the liquid flowing downwards
through the perforations. In current modern practice, as shown in the
adjacent diagram, better contacting is achieved by installing bubble-caps or
valve caps located at each perforation to promote the formation of vapor
bubbles flowing through a thin layer of liquid maintained by a weir on each
tray.
To design a distillation unit or a similar chemical process, the number of
theoretical trays or plates (that is, hypothetical equilibrium stages), N t,
required in the process should be determined, taking into account a likely
range of feedstock composition and the desired degree of separation of the
components in the output fractions. In industrial continuous fractionating
columns, N t is determined by starting at either the top or bottom of the
column and calculating material balances, heat balances and equilibrium
flash vaporizations for each of the succession of equilibrium stages until the
desired end product composition is achieved. The calculation process
requires the availability of a great deal of vapor-liquid equilibrium data for
the components present in the distillation feed, and the calculation procedure
is very complex.
In an industrial distillation column, the N t required to achieve a given
separation also depends upon the amount of reflux used. Using more reflux
decreases the number of plates required and using less reflux increases the
number of plates required. Hence, the calculation of N t is usually repeated
at various reflux rates. N t is then divided by the tray efficiency, E, to
determine the actual number of trays or physical plates, Na, needed in the
separating column. The final design choice of the number of trays to be
installed in an industrial distillation column is then selected based upon an
economic balance between the cost of additional trays and the cost of using a
higher reflux rate.
There is a very important distinction between the theoretical plate
terminology used in discussing conventional distillation trays and the
theoretical plate terminology used in the discussions below of packed bed
distillation or absorption or in chromatography or other applications. The
theoretical plate in conventional distillation trays has no "height". It is
simply a hypothetical equilibrium stage. However, the theoretical plate in
packed beds, chromatography and other applications is defined as having a
height.
Distillation and absorption packed beds

Distillation and absorption separation processes using packed beds for vapor
and liquid contacting have an equivalent concept referred to as the plate
height or the height equivalent to a theoretical plate (HETP). HETP arises
from the same concept of equilibrium stages as does the theoretical plate and
is numerically equal to the absorption bed length divided by the number of
theoretical plates in the absorption bed (and in practice is measured in this
way).

where:
Nt = the number of theoretical plates (also called the "plate count")
H = the total bed height
HETP = the height equivalent to a theoretical plate

The material in packed beds can either be random dumped packing (1-3" wide) such
as Raschig rings or structured sheet metal. Liquids tend to wet the surface of the
packing and the vapors contact the wetted surface, where mass transfer occurs.

Chromatographic Processes

The theoretical plate concept was also adapted


for chromatographic processes by Martin and Synge. The IUPAC's Gold
Book provides a definition of the number of theoretical plates in a
chromatography column.
The same equation applies in chromatography processes as for the packed
bed processes, namely:

where:
Nt = the number of theoretical plates (also called the "plate count")
H = the total column length
HETP = the height equivalent to a theoretical plate
Other Methods for Calculating No. of Trays
There are two different methods for calculating the no. of trays in
Distillation Column.

Fenske Equation
The Fenske equation in continuous fractional distillation is an equation used
for calculating the minimum number of theoretical plates required for the
separation of a binary feed stream by a fractionation column that is being
operated at total reflux (i.e., which means that no overhead product distillate
is being withdrawn from the column).
The equation was derived by Merrell Fenske in 1932, a professor who
served as the head of the chemical engineering department at
the Pennsylvania State University from 1959 to 1969.
This is one of the many different but equivalent versions of the Fenske
equation:

where:

N = minimum number of theoretical plates required at total reflux (of


which the reboiler is one)
Xd = mole fraction of more volatile component in the overhead distillate
Xb = mole fraction of more volatile component in the bottoms

αavg = average relative volatility of more volatile component to less


volatile component

For ease of expression, the more volatile and the less volatile components
are commonly referred to as the light key (LK) and the heavy key (HK),
respectively.
If the relative volatility of the light key to the heavy key is constant from
the column top to the column bottom, then αavg. is simply α. If the
relative volatility is not constant from top to bottom of the column, then the
following approximation may be used:
where:
αt = relative volatility of light key to heavy key at top of column
αb = relative volatility of light key to heavy key at bottom of column

The above Fenske equation can be modified for use in the total reflux
distillation of multi-component feeds.

Another form of the Fenske Equation


A derivation of another form of the Fenske equation for use in gas
chromatography is available on the U.S. Naval Academy's web
site. Using Raoult's law and Dalton's Law for a series of condensation and
evaporation cycles (i.e., equilibrium stages), the following form of the
Fenske equation is obtained:

where:
N = number of equilibrium stages
Zn = mole fraction of component n in the vapor phase
Xn = mole fraction of component n in the liquid phase
= vapor pressure of pure component n
McCabe-Thiele Method
The graphical approach presented by McCabe and Thiele in 1925,
the McCabe-Thiele method is considered the simplest and perhaps most
instructive method for analysis of binary distillation. This method uses the
fact that the composition at each theoretical tray (or equilibrium stage) is
completely determined by the mole fraction of one of the two components.
The McCabe-Thiele method is based on the assumption of constant molar
overflow which requires that:

• The molal heats of vaporization of the feed components are equal


• for every mole of liquid vaporized, a mole of vapour is condensed
• heat effects such as heats of solution and heat transfer to and from
the distillation column are negligible.

Construction and use of the McCabe-Thiele Diagram


Before starting the construction and use of a McCabe-Thiele diagram for the
distillation of a binary feed, the vapor-liquid equilibrium (VLE) data must be
obtained for the lower-boiling component of the feed.
Figure 1: Typical McCabe-Thiele diagram for distillation of a binary
feed
The first step is to draw equal sized vertical and horizontal axes of a graph.
The horizontal axis will be for the mole fraction (denoted by x) of the lower-
boiling feed component in the liquid phase. The vertical axis will be for the
mole fraction (denoted by y) of the lower-boiling feed component in the
vapor phase.
The next step is to draw a straight line from the origin of the graph to the
point where x and y both equal 1.0, which is the x = y line in Figure 1. Then
draw the equilibrium line using the VLE data points of the lower boiling
component, representing the equilibrium vapor phase compositions for each
value of liquid phase composition. Also draw vertical lines from the
horizontal axis up to the x = y line for the feed and for the desired
compositions of the top distillate product and the corresponding bottoms
product (shown in red in Figure 1).
The next step is to draw the operating line for the rectifying section (the
section above the feed inlet) of the distillation column, (shown in green in
Figure 1). Starting at the intersection of the distillate composition line and
the x = y line, draw the rectifying operating line at a downward slope
(Δy/Δx) of L / (D + L) where L is the molar flow rate of reflux and D is the
molar flow rate of the distillate product. For example, in Figure 1, assuming
the molar flow rate of the reflux L is 1000 moles per hour and the molar
flow rate of the distillate D is 590 moles per hour, then the downward slope
of the rectifying operating line is 1000 / (590 + 1000) = 0.63 which means
that the y-coordinate of any point on the line decreases 0.63 units for each
unit that the x-coordinate decreases.
Examples of q-line slopes
The next step is to draw the blue q-line (seen in Figure 1) from the x = y line
so that it intersects the rectifying operating line.
The parameter q is the mole fraction of liquid in the feed and the slope of the
q-line is q / (q - 1). For example, if the feed is a saturated liquid it has no
vapor, thus q = 1 and the slope of the q-line is infinite which means the line
is vertical. As another example, if the feed is all saturated vapor, q = 0 and
the slope of the q-line is 0 which means that the line is horizontal.
Some example q-line slopes are presented in Figure 2. As can be seen now,
the typical McCabe-Thiele diagram in Figure 1 uses a q-line representing a
partially vaporized feed.
Next, as shown in Figure 1, draw the purple operating line for the stripping
section of the distillation column (i.e., the section below the feed inlet).
Starting at the intersection of the red bottoms composition line and the x = y
line, draw the stripping section operating line up to the point where the blue
q-line intersects the green operating line of the rectifying section operating
line.
Finally, as exemplified in Figure 1, draw the steps between operating lines
and the equilibrium line and then count them. Those steps represent
the theoretical plates (or equilibrium stages). The required number of
theoretical plates is 6 for the binary distillation depicted in Figure 1.
Note that using colored lines is not required and only used here to make the
methodology easier to describe.
In continuous distillation with varying reflux ratio, the mole fraction of the
lighter component in the top part of the distillation column will decrease as
the reflux ratio decreases. Each new reflux ratio will alter the slope of the
rectifying section operating line.
When the assumption of constant molar overflow is not valid, the operating
lines will not be straight. Using mass and enthalpy balances in addition to
vapor-liquid equilibrium data and enthalpy-concentration data, operating
lines can be constructed based on Ponchon-Savarit's method.
Modelling
In the Mc Cabe-Thiele method we make the material balance and enthalpy
balance equation by dividing the distillation column into two two parts-the
enriching section and the stripping section .This are shown in the diagram as
well. The feed enters the column somewhere near the middle, overhead and
bottom products are withdrawn as shown. The column contains a number of
bubble cap plates. The vapour from the top plate passes to a condenser
where it is condensed to a saturated liquid.Some of the liquid from the
accumulator is returned as reflux continuously to the top plate.The residual
or the bottom product is taken out from the bottom as shown in the figure
below from an example.
Simulation
Computer simulation is the discipline of designing a model of an actual or
theoretical physical system, execution the model on a digital computer, and
analyzing the execution output. Simulation embodies the principle of
“learning by doing”---to learn about the system we must first build a model
of some sort and then operate the model. The use of simulation is an activity
that is as natural as a child who role plays. Children understand the world
around them by simulating (with toys and figurines) most of their
interactions with other people, animals and objects. As adults, we lose some
of this childlike behavior but recapture it later on through computer
simulation.
To understand the reality and all of its complexity, we must build artificial
objects and dynamically act out roles with them. Computer simulation is the
electronic equivalent of this type role playing and it serves to drive synthetic
environments and virtual world. Within the overall task of simulation, there
are three primary subfields: Model design, Model execution and Model
analysis.
To simulate something physical, you will first need to create a mathematical
model which represents that physical object. Models can take many forms
including declarative, functional, constraint, spatial or multimodal. A
multimodal is a model containing multiple integrated models each of which
represents a level of granularity for the physical system. The next task , once
a model has been developed, is to execute the model on a computer --- that
is, you need to create a program which steps through time while updating the
state and event variables in your mathematical model. There are many ways
to “step through time.” You can also execute the program on a massively
parallel computer. this is called parallel and distributed simulation.
The term simulation is used in different ways by different people. As used
here, simulation is defined as the process of creating a model (i.e., an
abstract representation or facsimile) of an existing or proposed system (e.g.,
a project, a business, a mine, a watershed, a forest, the organs in your body)
in order to identify and understand those factors which control the system
and/or to predict (forecast) the future behavior of the system.
The Power of Simulation
Simulation is a powerful and important tool because it provides a way in
which alternative designs, plans and/or policies can be evaluated without
having to experiment on a real system, which may be prohibitively costly,
time-consuming, or simply impractical to do. That is, it allows you to ask
“what if?” Question about a system without having to experiment on the
actual system itself (and hence in our the costs of field tests, prototypes, etc).

Why Do Simulation?
You may wonder whether simulation must be used to study dynamic
systems. There are many methods of modeling systems which involve the
solution of a closed form system. Simulation is often essential in the
following case:
1. The model is very complex with many variables and interacting
component;
2. The underlying relationships are nonlinear;
3. The model contain random varieties;
4. The model output is to be visual as in a 3D computer animation.
The power of simulation is that ---even for easily solvable linear systems---a
uniform model execution technique can be used to solve a large variety of
systems without resorting to a “bag of tricks” where one must choose
special-purpose and sometimes arcane solution methods to avoid simulation.

Types of Simulation Tools


The simulation tools are known as Simulator. The general purpose tools can
be broadly categorized as follows;

Discrete Event Simulators


These tools rely on a transaction-flow approach to modeling systems.
Models consist of entities, recourses, and control elements. Discrete
simulators are generally designed for simulating processes such as call
centers, factory operations, and shipping facilities in which the material or
information that is being simulated can be described as moving in discrete
steps or packets.
Agent based Simulator
This is a special class of discrete event simulator in which the mobile entities
are known as agent.

Continuous Simulators
This class of tools solves differential equations that describe the evolution of
a system using continuous equations. These types of simulators are the most
appropriate if the material or information that is being simulated can be
described as evolving or moving smoothly and continuously, rather than
infrequent discrete steps or packets.
A common class of continuous simulators is system dynamics tools based
on the standard stock and flow approach developed by Professor Jay
W.Forrester at MIT in the early 1960s.

Hybrid Simulator
These tools combine the features of continuous simulators and discrete
simulators. That is, they solve differential equations, but can superimpose
discrete events on the continuously varying system. Goldsim is a hybrid
simulator.
Problem
Included in table are the data for the McCabe-Thiele method at R=1.029, the
reflux ratio used for exact Ponchon-Savarit calculation. It is noteworthy that
for either method each at 1.5 times its respective value of Rm, the no. of
stages is essentially the same, and the maximum flow rates which would be
used to set the mechanical design of the tower are sufficiently similar for the
same final design to result.
Appendix

Simulation through C++

#include<iostream.h>
#include<math.h>
#include<conio.h>
# define R 8.314
# define t0 298
# define lpha 2.69
double pj_sat(double pj1,double pjx,double pjg1,double pjg2,double
pjp1,double pjp2);
double pj_satb(double pj1,double pjy,double pjg1,double pjg2,double
pjp1,double pjp2);
double bubl_t (double ax) ;
double dew_t(double ay,int flag );
double equili(double y);
double setq();
double liq(double th);
double vap(double th );
double inter();
double equi(double ax);
double stri(double x1);
double enri(double x);
class bubl { double t1,t2,t3,tn,y1,y2,temp;
double a1,a2,b1,b2,c1,c2;
public : double p1,x1,x2;
bubl()
{
a1=16.5938;
a2=16.262;
b1=3644.3;
b2=3799.89;
c1=239.76;
c2=226.35;
}
double anto(double p2sat,int flag) ;
double ant(double ta,int flag) ;
void read_data();
void show_data();
double solve_y();

};

double bubl::anto(double p2sat,int flag)


{double t1,t2;
if(flag==2)
{
t1=a2-log(p2sat);
t1=b2/t1;
t2=t1-c2;
return t2;
}
if(flag==1)
{
t1=a1-log(p2sat);
t1=b1/t1;
t2=t1-c1;
return t2;
}
}

double bubl::ant(double ta,int flag)


{ double t1;
if(flag==1)
{
t1= b1/(c1+ta);
t1=a1-t1;
t1=exp(t1);
return t1;
}
if(flag==2)
{
t1= b2/(c2+ta);
t1=a2-t1;
t1=exp(t1);
return t1;
}
}
void bubl::read_data()
{cout<<"ENTER THE VALUE OF PRESSURE \n ";
cin>> p1;
cout<<"ENTER THE VALUE OF MOLE FRACTION OF FIRST
COMPONENT X1 \n";
cin>>x1;
x2=1-x1;
}
/*double bubl::solve_y()
{ y1=(x1*g1*ant(t3,1))/(phi1*p1);
y2=1-y1;
}*/

class activity { double b12,b21,alpha,tou12,tou21,G12,G21;


public : double gamma1,gamma2;
activity()
{ b12=-253.88 ;
b21=845.21 ;
alpha=.2994 ;
}

void set_tou (double t );


void set_G ();
// void read_x ();
void set_gamma (double,double );
void show_gamma ();
};

void activity :: set_tou ( double t )


{
double t1,t2;
t2=t+273.15;
t1= 1.987*t2;
tou12=b12/t1;
tou21=b21/t1;
}
void activity :: set_G ()
{
G12=exp( -alpha*tou12);
G21=exp( -alpha*tou21);
}
/*void activity:: read_x ()
{ cout<<"ENTER THE VALUE OF MOLE FRACTION OF
COMPONENT 1\n";
cin>>x1;
cout<<"ENTER THE VALUE OF MOLE FRACTION OF
COMPONENT 2\n";
cin>>x2;
}*/
void activity:: set_gamma (double x1,double x2 )
{
double t1,t2,t3,t4;
t1=x1+x2*G21;
t2=x2+x1*G12;
t1=G21/t1;
t1=tou21*t1*t1;
t2=tou12/(t2*t2);
t2=G12*t2;
t1=t1+t2;
t1=x2*x2*t1 ;
gamma1=exp(t1);
t3=x1+x2*G21;
t3=tou21/(t3*t3);
t3=t3*G21;
t4=x2+x1*G12;
t4=G12/t4;
t4=tou12*t4*t4;
t4=t4+t3;
t4=x1*x1*t4;
gamma2=exp(t4);
}
double
xxd,xxf,xxw,rr,qq,n=0,ix,iy,xx2,t,p,x2,y2,conc_prof[15],temp_prof[20];
void activity:: show_gamma()
{cout<<"THE VALUE OF ACTIVITY COEFFICIENT FOR THE
FIRST COMPONENT IS \n";
cout<<gamma1<<"\n";
}

void main ()
{ // double xxd,xxf,xxw,rr,qq,n=0,ix,iy,xx2,t,p,x2,y2;
clrscr();
cout<<"ENTER THE MOLE FRACTION OF TOP PRODUCT: \n";
cin>>xxd;
cout<<"ENTER THE MOLE FRACTION OF BOTTOM PRODUCT:
\n";
cin>>xxw;
cout<<"ENTER THE MOLE FRACTION OF FEED: \n";
cin>>xxf;
cout<<"ENTER THE REFLUX RATIO: \n";
cin>>rr;
// cout<<"enter the operating conditions :"<<endl;
cout<<"ENTER THE TEMPERATURE OF FEED IN
CENTIGRADES:\n";
cin>>t;
cout<<"ENTER THE PRESSURE OF FEED IN kPa:\n";
cin>>p;
qq=setq();
ix=inter();
iy=enri(ix);
x2=equili(xxd);
for(;x2>xxw;n++)
{
if(n==0)
{
if(x2>=ix)
y2=enri(x2);
else
y2=stri(x2);
}
else
{
x2=equili(y2);
if(x2>=ix)
y2=enri(x2);
else
y2=stri(x2);
}
conc_prof[n]=x2;
temp_prof[n]=bubl_t(conc_prof[n]);
}
cout<<"THE THEORETICAL NUMBER OF PLATES ARE "<<endl;
cout<<n<<"\n";
for(int ii=0;ii<=n;ii++)
cout<<conc_prof[ii]<<"\n";
for( ii=0;ii<=n;ii++)
cout<<temp_prof[ii]<<"\n";

getch();
}
double bubl_t (double ax )
{
//clrscr();
double a1,a2,a3,a4,T;
double g1,g2;
bubl b;
activity a;
//b.read_data();/*READ VALUES OF PRESSURE AND X1*/
a1=b.anto(p,1);
a2=b.anto(p,2);
T=(a1*ax)+(a2*(1-ax));
while(1)
{
a1=b.ant(T,1);
a2=b.ant(T,2);
a3=T;
/* double temp;
cout<<"ENTER THE VALUE OF TEMPERATURE \n ";
cin>> temp;
activity a;*/
a.set_tou(T);
a.set_G();
// a.read_x();
a.set_gamma(ax,(1-ax));
g1=a.gamma1;
g2=a.gamma2;
a2=pj_sat(p,ax,g1,g2,a1,a2);/*SPECIES 2 IS J*/
T= b.anto(a2,2);
a4=(T-a3);
if( a4<0 )
a4=a4*-1;
if(a4 <.001)
break;
}
a3=(ax*g1*a1)/p;
return T;
}

double dew_t(double ay,int flag )


{
double a1,a2,a3,a4,T ;
double g1,g2,y1,y2;
bubl b;
activity a;
// cout<<"ENTER THE VALUE OF PRESSURE \n";
// cin>>b.p1;
// cout<<"ENTER THE VALUE OF MOLE FRACTION Y1 \n";
y1=ay;
y2=1-y1;
a1=b.anto(p,1);
a2=b.anto(p,2);
T=(a1*y1)+(a2*y2);
g1=1;
g2=1;
while(1)
{
a1=b.ant(T,1);
a2=b.ant(T,2);
a3=T;
b.x1=(y1*p)/(g1*a1);
b.x2=1-b.x1;
a.set_tou(T);
a.set_G();
// a.read_x();
a.set_gamma(b.x1,b.x2);
g1=a.gamma1;
g2=a.gamma2;
a1=pj_satb(p,y1,g1,g2,a1,a2);
T=b.anto(a1,1);
a4=(T-a3);
if(a4<0)
a4=a4*-1;
if(a4 <.001)
break;
}
if(flag==1)
return T;
else
return
b.x1;
}

double pj_sat(double pj1,double pjx,double pjg1,double pjg2,double


pjp1,double pjp2)
{
double z1,z2;
z1=pjx*pjg1*pjp1/pjp2;
z2=(1-pjx)*pjg2;
z2=pj1/(z1+z2);
return z2;
}

double pj_satb(double pj1,double pjy,double pjg1,double pjg2,double


pjp1,double pjp2)
{
double z1,z2;
z1=pjy/pjg1;
z2=((1-pjy)*pjp1)/(pjg2*pjp2);
z2=pj1*(z1+z2);
return z2;
}

double setq()
{
bubl b;
clrscr();
double hg,hl,hf,t1,t2,qe,fl,fg;
int ch;
cout<<"ENTER THE CONDITION OF FEED\n";
cout<<"1. Liquid \n";
cout<<"2. Saturated liquid\n";
cout<<"3. Mixture of saturated liquid & saturated vapor\n";
cout<<"4. Saturated vapor\n";
cout<<"5. Superheated vapor\n";
// cout<<"enter the feed condition : \n";
cin>>ch;
double bub_ret,dew_ret;
bub_ret= bubl_t(xxf);
hl=liq(bub_ret);
dew_ret=dew_t(xxf,1);
hg=vap(dew_ret);
switch(ch)
{
case 1:
hf=liq(t);
t1=hg-hf;
t2=hg-hl;
qe= t1/t2;
break;
case 2:
hf=hl;
qe=1;
break;
case 3:
cout<<"enter the total number of moles in the liquid feed :\n";
cin>>fl;
cout<<"enter the total number of moles in the vapor feed :\n";
cin>>fg;
hf=(hl*fl)+(hg*fg);
t1=hg-hf;
t2=hg-hl;
qe= t1/t2;
break;
case 4:
hf=hg;
qe=0;
break;
case 5:
hf=vap(t);
t1=hg-hf;
t2=hg-hl;
qe= t1/t2;
break;
};

return qe;
}
double liq(double th)
{
double t1, cl1=88.92;
// cl=(xxf*cl1)+((1-xxf)*cl2);
t1=th-t;
t1=cl1*t1;
return t1;
}
double vap(double th )
{
float t1,t2,t3,t4,v1=33536.268,v2=41157.68,y2 , cl1=87.18, cl2=75.44914;
t1=th-t;
t2=cl1*t1;
t3=cl2*t1;
t2=t2+v1;
t3=t3+v2;
t1=xxf*t2+(1-xxf)*t3;
return t1;

}
double inter()
{
double t1,t2,t3,t4;
t1=rr/(rr+1);
t2=xxf/(qq-1);
t3=qq/(qq-1);
t4=xxd/(rr+1);
return (t4+t2)/(t3-t1);
}
double equi(double ax )
{
double t1,t2,t3,g1;
bubl b;
activity a;
t1=bubl_t(xxd);
t3=b.ant(t1,1);
t2=dew_t(ax,2);
a.set_tou(t1);
a.set_G();
a.set_gamma(t2,t2-1);
g1=a.gamma1;
return (ax*p)/(g1*t3);
}
double enri(double x)
{
double t1,t2,t3;
t2=(rr+1);
t1=rr/t2;
t3=xxd/t2;
t1=(t1*x)+t3;
return t1;
}
double stri(double x1)
{
double s,t1,t2,t3,t4,i;
t1=iy-xxw;
t2=ix-xxw;
t3=-ix+x1;
t4=t1/t2;
s=t4;
t4=t4*t3;
t1=iy+t4;
i=s*ix;
i=-i+iy;
// cout<<"the slope"<<s<<endl;
// cout<<"the intercept"<<i;
return t1;
}

double equili(double y)
{
double a;
a=4.5;
double t1,t2;
t1=(a-(a*y) +y);
t2=y/t1;
return t2;
}
Bibliography

Mass-Transfer Operations-Robert E. Treybal

Unit Operations in Chemical Engg.-McCabe Smith Herriot

www.wikipedia.com

www.scienceworld.com

You might also like