Вы находитесь на странице: 1из 64

Drying

From Wikipedia, the free encyclopedia


Jump to navigationJump to search
Drying is a mass transfer process consisting of the removal of water or
another solvent[1] by evaporation from a solid, semi-solid or liquid. This process is often used as a
final production step before selling or packaging products. To be considered "dried", the final product
must be solid, in the form of a continuous sheet (e.g., paper), long pieces (e.g., wood), particles
(e.g., cereal grains or corn flakes) or powder (e.g., sand, salt, washing powder, milk powder). A
source of heat and an agent to remove the vapor produced by the process are often involved. In
bioproducts like food, grains, and pharmaceuticals like vaccines, the solvent to be removed is almost
invariably water. Desiccation may be synonymouswith drying or considered an extreme form of
drying.
In the most common case, a gas stream, e.g., air, applies the heat by convection and carries away
the vapor as humidity. Other possibilities are vacuum drying, where heat is supplied
by conduction or radiation (or microwaves), while the vapor thus produced is removed by
the vacuum system. Another indirect technique is drum drying (used, for instance, for manufacturing
potato flakes), where a heated surface is used to provide the energy, and aspirators draw the vapor
outside the room. In contrast, the mechanical extraction of the solvent, e.g., water,
by filtration or centrifugation, is not considered "drying" but rather "draining".

Contents

 1Drying mechanism

 2Methods of drying

 3Applications of drying

o 3.1Food

o 3.2Non-food products

o 3.3Sludges and fecal materials from sanitation processes

 4See also

 5Bibliography

 6References

 7External links

Drying mechanism[edit]
In some products having a relatively high initial moisture content, an initial linear reduction of the
average product moisture content as a function of time may be observed for a limited time, often
known as a "constant drying rate period". Usually, in this period, it is surface moisture outside
individual particles that is being removed. The drying rate during this period is mostly dependent on
the rate of heat transfer to the material being dried. Therefore, the maximum achievable drying rate
is considered to be heat-transfer limited. If drying is continued, the slope of the curve, the drying
rate, becomes less steep (falling rate period) and eventually tends to nearly horizontal at very long
times. The product moisture content is then constant at the "equilibrium moisture content", where it
is, in practice, in equilibrium with the dehydrating medium. In the falling-rate period, water migration
from the product interior to the surface is mostly by molecular diffusion, i,e. the water flux is
proportional to the moisture content gradient. This means that water moves from zones with higher
moisture content to zones with lower values, a phenomenon explained by the second law of
thermodynamics. If water removal is considerable, the products usually undergo shrinkage and
deformation, except in a well-designed freeze-drying process. The drying rate in the falling-rate
period is controlled by the rate of removal of moisture or solvent from the interior of the solid being
dried and is referred to as being "mass-transfer limited". This is widely noticed in hygroscopic
products such as fruits and vegetables, where drying occurs in the falling rate period with the
constant drying rate period said to be negligible. [2]

Methods of drying[edit]

In a typical phase diagram, the boundary between gas and liquid runs from the triple point to the critical point.
Regular drying is the green arrow, while supercritical drying is the red arrow and freeze drying is the blue.

The following are some general methods of drying:

 Application of hot air (convective or direct drying). Air heating increases the drying force for
heat transfer and accelerates drying. It also reduces air relative humidity, further increasing the
driving force for drying. In the falling rate period, as moisture content falls, the solids heat up and
the higher temperatures speed up diffusion of water from the interior of the solid to the surface.
However, product quality considerations limit the applicable rise to air temperature. Excessively
hot air can almost completely dehydrate the solid surface, so that its pores shrink and almost
close, leading to crust formation or "case hardening", which is usually undesirable. For instance
in wood (timber) drying, air is heated (which speeds up drying) though some steam is also
added to it (which hinders drying rate to a certain extent) in order to avoid excessive surface
dehydration and product deformation owing to high moisture gradients across timber
thickness. Spray drying belongs in this category.

 Indirect or contact drying (heating through a hot wall), as drum drying, vacuum drying. Again,
higher wall temperatures will speed up drying but this is limited by product degradation or case-
hardening. Drum drying belongs in this category.

 Dielectric drying (radiofrequency or microwaves being absorbed inside the material) is the
focus of intense research nowadays. It may be used to assist air drying or vacuum drying.
Researchers have found that microwave finish drying speeds up the otherwise very low drying
rate at the end of the classical drying methods.

 Freeze drying or lyophilization is a drying method where the solvent is frozen prior to drying
and is then sublimed, i.e., passed to the gas phase directly from the solid phase, below the
melting point of the solvent. It is increasingly applied to dry foods, beyond its already classical
pharmaceutical or medical applications. It keeps biological properties of proteins, and retains
vitamins and bioactive compounds. Pressure can be reduced by a high vacuum pump (though
freeze drying at atmospheric pressure is possible in dry air). If using a vacuum pump, the vapor
produced by sublimation is removed from the system by converting it into ice in a condenser,
operating at very low temperatures, outside the freeze drying chamber.

 Supercritical drying (superheated steam drying) involves steam drying of products containing
water. This process is feasible because water in the product is boiled off, and joined with the
drying medium, increasing its flow. It is usually employed in closed circuit and allows a
proportion of latent heat to be recovered by recompression, a feature which is not possible with
conventional air drying, for instance. The process has potential for use in foods if carried out at
reduced pressure, to lower the boiling point.

 Natural air drying takes place when materials are dried with unheated forced air, taking
advantage of its natural drying potential. The process is slow and weather-dependent, so a wise
strategy "fan off-fan on" must be devised considering the following conditions: Air temperature,
relative humidity and moisture content and temperature of the material being dried. Grains are
increasingly dried with this technique, and the total time (including fan off and on periods) may
last from one week to various months, if a winter rest can be tolerated in cold areas.

Applications of drying[edit]

Drying of fish in Lofoten in the production of stockfish

Food[edit]
Main article: Drying (food)
Foods are dried to inhibit microbial development and quality decay. However, the extent of drying
depends on product end-use. Cereals and oilseeds are dried after harvest to the moisture content
that allows microbial stability during storage. Vegetables are blanched before drying to avoid rapid
darkening, and drying is not only carried out to inhibit microbial growth, but also to avoid browning
during storage. Concerning dried fruits, the reduction of moisture acts in combination with its acid
and sugar contents to provide protection against microbial growth. Products such as milk powder
must be dried to very low moisture contents in order to ensure flowability and avoid caking. This
moisture is lower than that required to ensure inhibition to microbial development. Other products as
crackers are dried beyond the microbial growth threshold to confer a crispy texture, which is liked by
consumers.
Non-food products[edit]
Among non-food products, some of those that require considerable drying are wood (as part of
timber processing), paper, flax, and washing powder. The first two, owing to their organic origins,
may develop mold if insufficiently dried. Another benefit of drying is a reduction in volume and
weight.
Sludges and fecal materials from sanitation processes [edit]
In the area of sanitation, drying of sewage sludge from sewage treatment plants, fecal
sludge or feces collected in urine-diverting dry toilets (UDDT) is a common method to
achieve pathogen kill, as pathogens can only tolerate a certain dryness level. In addition, drying is
required as a process step if the excreta based materials are meant to be incinerated. [3]
The

SOLIDS DRYING: BASICS AND


APPLICATIONS
By Dilip M. Parikh, DPharma Group Inc. | April 1, 2014
50

Adjustment and control of moisture levels in solid materials through drying is a critical process in the manufacture
of many types of chemical products. As a unit operation, drying solid materials is one of the most common and
important in the chemical process industries (CPI), since it is used in practically every plant and facility that
manufactures or handles solid materials, in the form of powders and granules.

The effectiveness of drying processes can have a large impact on product quality and process efficiency in the CPI.
For example, in the pharmaceutical industry, where drying normally occurs as a batch process, drying is a key
manufacturing step. The drying process can impact subsequent manufacturing steps, including tableting or
encapsulation and can influence critical quality attributes of the final dosage form.

Apart from the obvious requirement of drying solids for a subsequent operation, drying may also be carried out to
improve handling characteristics, as in bulk powder filling and other operations involving powder flow; and to
stabilize moisture-sensitive materials, such as pharmaceuticals.

This article provides basic information on the sometimes complicated heat- and mass-transfer processes that are
important in drying, and discusses several technologies used to accomplish the task.

MECHANISM OF DRYING
Drying may be defined as the vaporization and removal of water or other liquids from a solution, suspension, or
other solid-liquid mixture to form a dry solid. It is a complicated process that involves simultaneous heat and mass
transfer, accompanied by physicochemical transformations. Drying occurs as a result of the vaporization of liquid by
supplying heat to wet feedstock, granules, filter cakes and so on. Based on the mechanism of heat transfer that is
employed, drying is categorized into direct (convection), indirect or contact (conduction), radiant (radiation) and
dielectric or microwave (radio frequency) drying.

Heat transfer and mass transfer are critical aspects in drying processes. Heat is transferred to the product to
evaporate liquid, and mass is transferred as a vapor into the surrounding gas. The drying rate is determined by the
set of factors that affect heat and mass transfer. Solids drying is generally understood to follow two distinct drying
zones, known as the constant-rate period and the falling-rate period. The two zones are demarcated by a break point
called the critical moisture content.

In a typical graph of moisture content versus drying rate and moisture content versus time (Figure 1), section AB
represents the constant-rate period. In that zone, moisture is considered to be evaporating from a saturated surface at
a rate governed by diffusion from the surface through the stationary air film that is in contact with it. This period
depends on the air temperature, humidity and speed of moisture to the surface, which in turn determine the
temperature of the saturated surface. During the constant rate period, liquid must be transported to the surface at a
rate sufficient to maintain saturation.

Figure 1. Segment AB of the graph represents the constant-rate drying period, while segment BC is the falling-rate
period

At the end of the constant rate period, (point B, Figure 1), a break in the drying curve occurs. This point is called the
critical moisture content, and a linear fall in the drying rate occurs with further drying. This section, segment BC, is
called the first falling-rate period. As drying proceeds, moisture reaches the surface at a decreasing rate and the
mechanism that controls its transfer will influence the rate of drying. Since the surface is no longer saturated, it will
tend to rise above the wet bulb temperature. This section, represented by segment CD in Figure 1 is called the
second falling-rate period, and is controlled by vapor diffusion. Movement of liquid may occur by diffusion under
the concentration gradient created by the depletion of water at the surface. The gradient can be caused by
evaporation, or as a result of capillary forces, or through a cycle of vaporization and condensation, or by osmotic
effects.

The capacity of the air (gas) stream to absorb and carry away moisture determines the drying rate and establishes the
duration of the drying cycle. The two elements essential to this process are inlet air temperature and air flowrate.
The higher the temperature of the drying air, the greater its vapor holding capacity. Since the temperature of the wet
granules in a hot gas depends on the rate of evaporation, the key to analyzing the drying process is psychrometry,
defined as the study of the relationships between the material and energy balances of water vapor and air mixture.

DRYING ENDPOINT
There are a number of approaches to determine the end of the drying process. The most common one is to construct
a drying curve by taking samples during different stages of drying cycle against the drying time and establish a
drying curve. When the drying is complete, the product temperature will start to increase, indicating the completion
of drying at a specific, desired product-moisture content. Karl Fischer titration and loss on drying (LOD) moisture
analyzers are also routinely used in batch processes. The water vapor sorption isotherms are measured using a
gravimetric moisture-sorption apparatus with vacuum-drying capability.

For measuring moisture content in grain, wood, food, textiles, pulp, paper, chemicals, mortar, soil, coffee, jute,
tobacco, rice and concrete, electrical-resistance-type meters are used. This type of instrument operates on the
principle of electrical resistance, which varies minutely in accordance with the moisture content of the item
measured. Dielectric moisture meters are also used. They rely on surface contact with a flat plate electrode that does
not penetrate the product.

For measuring moisture content in paper rolls or stacks of paper, advanced methods include the use of the radio
frequency (RF) capacitance method. This type of instrument measures the loss, or change, in RF dielectric constant,
which is affected by the presence or absence of moisture.

TYPES OF DRYERS
Adiabatic dryers are the type where the solids are dried by direct contact with gases, usually forced air. With these
dryers, moisture is on the surface of the solid. Non-adiabatic dryers involve situations where a dryer does not use
heated air or other gases to provide the energy required for the drying process

Dryer classification can also be based on the mechanisms of heat transfer as follows:

• Direct (convection)

• Indirect or contact (conduction)

• Radiant (radiation)

• Dielectric or microwave (radio frequency) drying

Direct, or adiabatic, units use the sensible heat of the fluid that contacts the solid to provide the heat of vaporization
of the liquid.

With adiabatic dryers, solid materials can be exposed to the heated gases via various methods, including the
following:

• Gases can be blown across the surface (cross circulation)

• Gases can be blown through a bed of solids (through-circulation); used when solids are stationary, such as wood,
corn and others

• Solids can be dropped slowly through a slow-moving gas stream, as in a rotary dryer
• Gases can be blown through a bed of solids that fluidize the particles. In this case, the solids are moving, as in a
fluidized-bed dryer

• Solids can enter a high-velocity hot gas stream and can be conveyed pneumatically to a collector (flash dryer)

Non-adiabatic dryers (contact dryers) involve an indirect method of removal of a liquid phase from the solid
material through the application of heat, such that the heat-transfer medium is separated from the product to be dried
by a metal wall. Heat transfer to the product is predominantly by conduction through the metal wall and the impeller.
Therefore, these units are also called conductive dryers.

Although more than 85% of the industrial dryers are of the convective type, contact dryers offer higher thermal
efficiency and have economic and environmental advantages over convective dryers. Table 1 compares direct and
indirect dryers, while Table 2 shows the classification of dryers based on various criteria.

Table 1. Comparison of Direct and Indirect dryers [4]


Direct/adiabatic dryer (convective Indirect/non-adiabatic contact
Property
type) dryer (conductive type)
Uses sensible heat of gas that contacts Little or no carrier gas is required
Carrier gas the solid to provide the heat of to remove the vapors released from
vaporization of the liquid the solids
Heat transfer medium is in direct contact
Heat needed to vaporize the solvent
Heat transfer with the surface of the material to be
is transferred through a wall
dried
Avoided, as the heat transfer
Risk of cross
Persists medium does not contact the
contamination
product
Difficult as there is a large volume of Easier because of limited amount of
Solvent recovery
gas to be cooled to recover the solvent non-condensable gas encountered
Operation under Allows operation under vacuum,
Not possible
vacuum ideal for heat sensitive materials
Minimized because of small
Dusting High
volume of vapors involved
Easier to control as vapors can be
Explosion hazard Higher rate
easily condensed
Handling of toxic
Not suitable Suitable because of low gas flow
materials
Higher energy efficiency as the
Significant energy lost through exhaust
Energy efficiency energy lost through the exhaust gas
gas
is greatly reduced
Drying rates are limited by heat
Evaporation and
Higher than contact dryers transfer area, lower production
production rates
rates
Higher initial cost; difficult to
Cost High
design, fabricate and maintain
Table 2. Classification of Dryers [5]
Criterion Types
Mode of operation Batch Continuous*
Convection*, conduction, radiation, electromagnetic fields,
Heat input type combination of heat transfer modes Intermittent or continuous*
Adiabatic or non-adiabatic
State of material in dryer Stationary Moving agitated, dispersed
Operating pressure Vacuum* Atmospheric
Drying medium
Air* Superheated steam Fluegases
(convection)
Below boling temperature* Above boiling temperature Below
Drying temperature
freezing point
Relative motion between Co-current Countercurrent Mixed flow
drying medium and solids
Number of stages Single* Multistage
Residence time Short (60 min)
* Most common in practice

BATCH DRYERS
The following are descriptions of various types of batch dryers.

Tray dryers. This dryer type operates by passing hot air over the surface of a wet solid that is spread over trays
arranged in racks. Tray dryers are the simplest and least-expensive dryer type. This type is most widely used in the
food and pharmaceutical industries. The chief advantage of tray dryers, apart from their low initial cost, is their
versatility. With the exception of dusty solids, materials of almost any other physical form may be dried. Drying
times are typically long (usually 12 to 48 h).

Vacuum dryers. Vacuum dryers offer low-temperature drying of thermolabile materials or the recovery of solvents
from a bed. Heat is usually supplied by passing steam or hot water through hollow shelves. Drying temperatures can
be carefully controlled and, for the major part of the drying cycle, the solid material remains at the boiling point of
the wetting substance. Drying times are typically long (usually 12 to 48 h).

For more on vacuum drying, see Batch Drying with Vacuum Contact
Dryershttps://www.chemengonline.com/batch-drying-with-vacuum-contact-dryers-2/

Fluidized-bed dryers. A gas-fluidized bed may have the appearance of a boiling liquid. It has bubbles, which rise and
appear to burst. The bubbles result in vigorous mixing. A preheated stream of air enters from the bottom of the
product container holding the product to be dried and fluidizes it. The resultant mixture of solids and gas behave like
a liquid, and thus the solids are said to be fluidized. The solid particles are continually caught up in eddies and fall
back in a random boiling motion so that each fluidized particle is surrounded by the gas stream for efficient drying,
granulation or coating purposes. In the process of fluidization, intense mixing occurs between the solids and air,
resulting in uniform conditions of temperature, composition and particle size distribution throughout the bed.

Freeze dryers. Freeze-drying is an extreme form of vacuum drying in which the water or other solvent is frozen and
drying takes place by subliming the solid phase. Freeze-drying is extensively used in two situations: (1) when high
rates of decomposition occur during normal drying; and (2) with substances that can be dried at higher temperatures,
and that are thereby changed in some way.

Microwave vacuum dryers. High-frequency radio waves with frequencies from 300 to 30,000 MHz are utilized in
microwave drying (2,450 MHz is used in batch microwave processes). Combined microwave-convective drying has
been used for a range of applications at both laboratory and industrial scales. The bulk heating effect of microwave
radiation causes the solvent to vaporize in the pores of the material. Mass transfer is predominantly due to a pressure
gradient established within the sample. The temperature of the solvent component is elevated above the air
temperature by the microwave heat input, but at a low level, such that convective and evaporative cooling effects
keep the equilibrium temperature below saturation. Such a drying regime is of particular interest for drying
temperature-sensitive materials. Microwave-convective processing typically facilitates a 50% reduction in drying
time, compared to vacuum drying.

CONTINUOUS DRYERS
Continuous dryers are mainly used in chemical and food industries, due to the large volume of product that needs to
be processed. Most common are continuous fluid-bed dryers and spray dryers. There are other dryers, depending on
the product, that can be used in certain industries — for example, rotary dryers, drum dryers, kiln dryers, flash
dryers, tunnel dryers and so on. Spray dryers are the most widely used in chemical, dairy, agrochemical, ceramic and
pharmaceutical industries.

Spray dryer. The spray-drying process can be divided into four sections: atomization of the fluid, mixing of the
droplets, drying, and, removal and collection of the dry particles (Figure 2). Atomization may be achieved by means
of single-fluid or two-fluid nozzles, or by spinning-disk atomizers. The flow of the drying gas may be concurrent or
countercurrent with respect to the movement of droplets. Good mixing of droplets and gas occurs, and the heat- and
mass-transfer rates are high. In conjunction with the large interfacial area conferred by atomization, these factors
give rise to very high evaporation rates. The residence time of a droplet in the dryer is only a few seconds (5–30 s).
Since the material is at wet-bulb temperature for much of this time, high gas temperatures of 1,508 to 2,008C may
be used, even with thermolabile materials. For these reasons, it is possible to dry complex vegetable extracts, such as
coffee or digitalis, milk products, and other labile materials without significant loss of potency or flavor. The capital
and running costs of spray dryers are high, but if the scale is sufficiently large, they may provide the cheapest
method.

For more on spray drying see Optimizing Analysis for Spray-


Drying https://www.chemengonline.com/optimizing-analysis-for-spray-drying/

Figure 2. Spray dryers have high capital and operating costs, but can be the least expensive method at large scales
Source: Anhydro

DRYER EFFICIENCY
With increasing concern about environmental degradation, it is desirable to decrease energy consumption in all
sectors. Drying has been reported to account for anywhere from 12 to 20% of the energy consumption in the
industrial sector. Drying processes are one of the most energy-intensive unit operations in the CPI.

One measure of efficiency is the ratio of the minimum quantity of heat that will remove the required water to the
energy actually provided for the process. Sensible heat can also be added to the minimum, as this added heat in the
material often cannot be economically recovered. Other newer technologies have been developed, such as sonic
drying, superheated steam, heat-pump-assisted drying and others.

CONCLUDING REMARKS
Drying is an essential unit operation used in various process industries. The mechanism of drying is well understood
as a two-stage process and depends on the drying medium and the moisture content of the product being dried.

Batch dryers are common in chemical and pharmaceutical industries, while continuous dryers are routinely used
where large production is required. Since the cost of drying is a significant portion of the cost of manufacturing a
product, improving efficiency or finding alternative drying routes is essential.

Edited by Scott Jenkins


FURTHER READING
1. Séverine, Thérèse, Mortier, F.C., De Beer, Thomas, Gernaey, Krist V., Vercruysse, Jurgen, et al. “Mechanistic
modelling of the drying behavior of single pharmaceutical granules,” European Journal of Pharmaceutics and
Biopharmaceutics 80, pp. 682–689, 2012.

2. Mezhericher, M., Levy, A. and Borde, I., “Theoretical drying model of single droplets containing insoluble or
dissolved solids,” Dry. Technol. 25 (6), pp. 1025– 1032, 2007.

3. Mezhericher, M., Levy, A. and Borde, I., “Modelling of particle breakage during drying,” Chem. Eng.
Progress. 47(8), pp. 1404–1411, 2008.

4. Sahnia, E.K., Chaudhuria, B., “Contact drying: A review of experimental and mechanistic modeling
approaches,” International Journal of Pharmaceutics, 434 pp. 334–348, 2012.

5. Mujumdar, A., “Handbook of Industrial Drying” 2nd ed. edited by Mujumdar, Marcel Dekker Publishing, 1995.

6. Raghavan, G.S.V., Rennie, T.J., Sunjka, P.S., Orsat, V., Phaphuangwittayakul, W. and Terdtoon, P., “Overview of
new techniques for drying biological materials, with emphasis on energy aspects,” Brazilian Journal of Chemical
Engineering, 22(2), pp. 195–201, 2005.

Source: Michael G. Benton and Kerry M. Dooley, Department of Chemical Engineering,


Louisiana State University, Baton Rouge, LA

Dryers are utilized in numerous industrial processes. The function of a dryer is to use
heat transfer processes to dry solids. A variety of dryer types exist. Adiabatic dryers use
convection and direct contact with gases to dry solids, whereas non-adiabatic dryers
use methods other than heated gas contact to dry 1, including conduction, radiation, and
radio frequency drying1. Dryers can be used for batch processes or they be in
continuous use1.

In this experiment, the effects of temperature and air velocity on the drying rate of sand
will be determined using a tray dryer. Three different power settings (1000 W, 1500 W
and 2500 W) for two different air flow rates will be tested, providing a total of six data
sets. From this data, the heat and mass transfer coefficients can be calculated.

CITE THIS VIDEO

JoVE Science Education Database. Chemical Engineering. Using a Tray Dryer to


Investigate Convective and Conductive Heat Transfer. JoVE, Cambridge, MA, (2018).

PRINCIPLES

Tray dryers are one type of batch dryer, which also include fluidized-bed dryers, freeze
dryers and vacuum dryers. Tray dryers use convective heat transfer to flow heated air
over solids to dry them. They are used by a variety of industries including for the
production of pharmaceuticals and other chemicals 1. Continuous dryers on the other
hand are common to large volume product industries, such as the food industry 1.

To begin the process in a typical tray dryer, the tray is filled evenly with a wet solid, such
as sand, and loaded into the apparatus. The dryer’s adjustable fan and heater allow for
continuous variations in flow rate from the fan through the drying channel, and heat duty
variations in 500 watt increments. As the dryer operates, water evaporates from the
sand into the air. The drying rate is then calculated by weighing the initial solid/water
mixture and subtracting the weight of the final dry solid and at various timed intervals.

Heat transfer is driven by the temperature difference between the sand and the
surrounding air. A simplified Newton’s Law of Heating (Equation 1) can be used to
model the heat transfer between the heated air and sand-air interface to obtain an
experimental heat transfer coefficient. Other heat duty terms are negligible compared to
the term in Equation 1,

Equation 1

where q is the heat transferred, ṁ is the water evaporated in an allotted amount of time
or rate of evaporation, ∆Hvapis the enthalpy of vaporization, hy is the heat transfer
coefficient, Tair is the air temperature, and Ts is the sand’s surface temperature.

In order to obtain an experimental mass transfer coefficient, the transfer of water from
sand to air will be modeled as mass transfer flowing across a true phase boundary. The
drying rate equation (Equation 2) is this model.

Equation 2

where ky is the mass transfer coefficient, C is the concentration of water, and A is the
surface area of the boundary. Concentrations of water in the sand (C s) and air (C∞) will
be obtained by using a mass balance and psychrometric charts, respectively. These are
used to solve for the drying rate.

Theoretical values can be compared to the experimental data by calculating heat and
mass transfer coefficients. The theoretical heat (Equation 3) and mass (Equation 4)
transfer coefficients are obtained from the properties of the substances involved from
correlations.

Equation 3

Equation 4

where Re is the Reynolds number, Pr is the Prandtl number, Sc is the Schmidt number,
DAB is the diffusivity of water in air, and L is the length.

PROCEDURE

The experiment will consist of four runs, each testing a different combination of one of
two fan and heat settings.

1. Tray Dryer Operation

1. Prepare the slurry by mixing 500 g of sand and 150 mL of water and load into the
experimental tray for the unit. Spread the mixture evenly in the tray.

2. With the unit off, place the tray in the drying chamber.

3. Turn on the main unit, then turn on the blower and heater.

4. Set the air velocity and temperature for each run. The three air velocities should
range from 0.8 ft/s to 2.0 ft/s (one high, medium, and low) with a constant temperature
around 195 ºF. The three temperatures should range from 130 - 200 ºF with a constant
air velocity of 1.8 ft/s.

5. Take measurements every 5 min over the entire run, which should last 45 min.
The data collected should include inlet air temperature, sand temperature, sand weight,
outlet air temperature, outlet air flow, dry bulb temperature, and wet bulb temperature.
Use the digital thermometers for temperature readings, air flow settings for the air flow,
a digital scale for sand weight and a sling psychrometer for the wet and dry bulb
temperatures.

6. Repeat the process for each set of settings, totaling four unique runs.

RESULTS

From the data collected, the following information can be obtained. Use psychrometric
charts to determine the absolute humidity, which gives the concentration of water
present in the air. The heat transfer coefficients can be calculated using the measured
temperatures and Equation 1. And finally, the change in mass of the wet sand can be
used to calculate the concentration of water in the sand.

The moisture content of sand decreased linearly over time. As expected, the
evaporation rate was found to increase with larger flow rate and heat duty. According to
their equations, both heat and mass transfer coefficients are directly proportional to the
evaporation rate at the sand-air interface. Theoretical values of heat and mass transfer
coefficients were found to have a strong positive correlation with a R2 of 99%. The
experimental values only showed a weak correlation after testing.

The relationships between air flow and evaporation rate and between temperature and
evaporation rate both increased linearly (Figure 1, Figure 2). Increased air flow (Figure
1) and increased temperature (Figure 2) both increased the evaporation rate. These
graphs show that when air flow or temperature increase and the other variable is held
constant, the evaporation rate will increase at an equivalent rate and follow a positive
linear trend. The air flow variation test was a measure of convective heat transfer, while
the temperature variation test was a measure of conductive heat transfer. The sum of
the two tests shows that both convective and conductive heat transfer follow a linear
relationship with evaporation rate.

Figure 1: Depiction of the relationship between air velocity and evaporation rate,
which increased linearly.
Figure 2: Depiction of the relationship between temperature and evaporation rate,
which increased linearly.

There are many sources of error in the measurements with the greatest sources for
error being the relative humidity and temperature of the air-sand interface. Also, air flow
rate effect on the weight of the tray was deemed unimportant but it is a source of error.
Some of this error may have also reduced the correlation of the heat and mass transfer
coefficients. These coefficients were calculated theoretically and proven to be
correlated. However, the experimental data did not show a significant trend, despite
being theoretically similar.

APPLICATIONS AND SUMMARY

A tray dryer was used to measure the drying rate of sand with respect to convective and
conductive heat transfer. Using the dryer at three different power levels and two different
flow rates, six experimental data sets were found. Measurements were taken by
weighing the sand/water mixture at five minute intervals.

This experiment made use of Newton’s Law of Heating, drying rate modeling, and heat
and mass transfer modeling. Heat and mass transfer coefficients were determined with
the use of a boundary layer model. Theoretically, the heat and mass transfer coefficients
show a very strong positive linear correlation. Even though the experimental results
showed a positive trend as well, the data was too inaccurate to display any significant
correlation between the two.

Tray-drying can be used in a variety of fields. One such field is pharmaceuticals. In


pharmaceuticals, tray dryers are used to dry many different base materials, including
sticky, granular, and crystalline materials2. Many plastics used in pharmaceuticals can
be dried in tray dryers2. Additionally, precipitates, pastes, and other wet masses can be
dried with a tray dryer, along with crude drugs, chemicals, powders, and tablet granules.
Even some equipment is dried in the dryers2. Tray dryers offer many advantages to this
industry, since they are used for batches, which can vary in size and be handled without
losses2. The dryers are also readily adjusted to accompany other materials in an
efficient manner2. In some cases, tray dryers in a vacuum are used to dry heat sensitive
products like vitamins2.

Tray dryers are also used in food processing3. Food can be spread out thinly and evenly
onto the trays for drying3. Depending on the type of food, drying can be performed by
heating with air moving across the trays, conduction from heated trays or shelves, or
radiation form other heated surfaces3. Air can be used with the additional benefit of
removing moist vapors, though this can be a problem for some foods 3.
REFERENCES

1. "Solids Drying: Basics and Applications - Chemical Engineering." Chemical


Engineering Solids Drying Basics and Applications Comments. N.p., n.d. Web. 12 Jan.
2017.

2. "Pharmainfo.net." Tray dryer by Saraswathi.B. N.p., n.d. Web. 12 Jan. 2017.

3. "Unit Operations in Food Processing - R. L. Earle." Unit Operations in Food


Processing - R. L. Earle. N.p., n.d. Web. 12 Jan. 2017.

Laboratory 6.2: Distillation:


Purify Ethanol
This article incorporates, in modified form,
material from Illustrated Guide to Home
Chemistry Experiments: All Lab, No Lecture.

Distillation is the oldest method used for


separating mixtures of liquids. Distillation
exploits the fact that different liquids have
different boiling points. When a mixture of liquids
is heated, the liquid with the lower (or lowest)
boiling point vaporizes first. That vapor is routed through a condenser,
which cools the vapor and causes it to condense as a liquid; the liquid
is then collected in a receiving vessel. As the original liquid mixture
continues being heated, eventually, some or all of the lower-boiling
liquid is driven off, leaving only the higher-boiling liquid or liquids in
the distillation vessel.

I say “some or all” because distillation is an imperfect method for


separating mixtures of liquids that form azeotropes. An azeotrope, also
called a constant boiling mixture, is a mixture of two or more liquids at
a specific ratio, whose composition cannot be altered by simple
distillation. Every azeotrope has a characteristic boiling point, which
may be lower (a positive azeotrope or minimum-boiling mixture) or
higher (a negative azeotrope or maximum-boiling mixture) than the
boiling points of the individual liquids that make up the azeotrope.

For example, ethanol forms a positive azeotrope with water. The boiling
point of a mixture of 95.6% ethanol (by weight) with 4.4% water is 78.1
°C, which is lower than the boiling point of pure water (100 °C) or pure
ethanol (78.4 °C). Because the azeotropic mixture boils at a lower
temperature, it’s impossible to use simple distillation to produce
ethanol at concentrations higher than 95.6%. (More concentrated
ethanol solutions can be produced by using drying agents such as
anhydrous calcium chloride that physically absorb the water from a
95.6% solution of ethanol. These solutions must be stored and handled
carefully, because otherwise they absorb water vapor from the air until
they reach the 95.6% azeotropic concentration.)

Ethanol also forms azeotropes with many other liquids, including some
that are poisonous or taste bad. This allows production of denatured
ethanol, which is toxic, cannot be drunk and so can be sold cheaply
without cannibalizing sales of (and taxes on) much more expensive
potable ethanol, such as vodka and other distilled beverages.
High Boiling Point Azeotropes
Hydrochloric acid is one familiar example of a negative (high-boiling)
azeotrope. Pure hydrogen chloride has a boiling point of -84 °C and
pure water a boiling point of 100 °C. A solution of 20.2% hydrogen
chloride (by weight) in water has a boiling point of 110 °C, higher than
the boiling point of either component. This means that boiling a
solution of hydrochloric acid of any concentration eventually produces
a solution of exactly 20.2% hydrogen chloride by weight. If the starting
solution is more dilute, water is driven off until the solution reaches
20.2% concentration. If the starting solution is more concentrated,
hydrogen chloride gas is driven off until the solution reaches 20.2%
concentration.
In this laboratory, we’ll use distillation to increase the concentration of
an ethanol solution. At 25 °C, pure water has a density of 0.99704 g/mL
and pure ethanol a density of 0.78522 g/mL. Solutions of ethanol and
water have densities between these figures. If you add the densities of
the pure liquids and divide by two, you get 0.89113 g/mL, which you
might assume is the density of a 50/50 ethanol-water mixture. As it
turns out, that’s not true. Ethanol and water do not mix volumetrically;
that is, if you mix 100 mL of pure ethanol with 100 mL of pure water,
you do not get 200 mL of solution, for the same reason that dissolving
100 mL of sucrose in 100 mL of water does not yield 200 mL of
solution.
Nonetheless, it’s possible to determine the concentration of ethanol by
measuring the density of the solution and comparing that value to
ethanol-water density tables available in the CRC handbook and similar
publications. We’ll measure the density of the starting solution and the
resulting distillate and compare those values to published values to
determine the ethanol concentrations of the two solutions.
Required Equipment and Supplies

 goggles, gloves, and protective clothing

 balance (optional, but strongly recommended)

 cylinder, graduated, 100 mL


 flask, Erlenmeyer, 250 mL

 thermometer (optional)

 stopper, 2-hole (to fit Erlenmeyer flask)

 glass tubing, 75 mm (2)

 flexible tubing, ~ 50 cm

 pipette, disposable

 hotplate or kitchen stove burner

 coffee can, pickle jar, or similar tall wide-mouth container

 glycerol (see Substitutions and Modifications)

 ethanol, 70% (100 mL)

 tap water

 ice

All of the specialty lab equipment and chemicals needed for this and
other
lab sessions are available individually from Maker Shed or other
laboratory
supplies vendors. Maker Shed also offers customized laboratory kits at
special
prices, including the Basic Laboratory Equipment Kit, the Laboratory
Hardware Kit, the Volumetric Glassware Kit, the Core Chemicals Kit,
and the
Supplemental Chemicals Kit.
CAUTION

Ethanol solutions of 45% or greater concentration are


flammable. Ethanol vapor is extremely flammable. Handle the
hot liquids used in this experiment with extreme care and
have a fire extinguisher handy. The distillate produced in this
experiment is not safe to drink. Ethanol is denatured with compounds
that form azeotropes with ethanol, specifically to ensure that
denatured ethanol cannot be undenatured by distillation. Wear splash
goggles, gloves, and protective clothing.
Substitutions and Modifications
 If you do not have a balance, you can still do this
experiment to observe the distillation process, but you will
not be able to determine the ethanol concentration of the
distillate.

 If you do not have a thermometer, plug the second hole in


the stopper with a small piece of duct tape or simply by
dropping a disposable pipette into the hole. (Not much
pressure is generated because the ethanol vapor can
escape from the flask through the glass and flexible tubing.)

 If you do not have glycerol (sold in drugstores under that


name, or as glycerin), you can use vegetable oil or mineral
oil to lubricate the stopper before you insert the
thermometer and glass tubing. If you use oil, wash the
stopper thoroughly with soap or detergent as soon as
possible after you complete the experiment. Allowing any
type of oil to remain in contact with the stopper for too long
can damage it.

 Ethanol is sold in drugstores, sometimes by name, and


sometimes as ethyl alcohol, ethyl rubbing alcohol, or simply
rubbing alcohol. (If the bottle is labeled as rubbing alcohol,
make sure it’s ethanol rather than isopropanol.) Most
drugstore ethanol is 70% by volume (about 64.7% by weight),
but 90% and 95% concentrations are also common. If you
start with more concentrated ethanol, dilute it with water to
70% before you begin the distillation.

Figure 6-2 shows the distillation apparatus I originally used for this lab
session. Note that both flasks are securely clamped to prevent tipping,
and that the stopper in the receiving flask (right) has an open hole to
vent the flask and prevent pressure buildup. The receiving flask sits in
a large container of ice water. As the alcohol vapor passes through the
tubing and into the receiving flask, it condenses immediately when it
contacts the cold air in the receiving flask. For this modified version of
the lab session, we’ll use a simplified distillation apparatus,
substituting a 100 mL graduated cylinder for the second Erlenmeyer
flask as the receiving vessel. We’ll also use a thermometer in the
distillation flask to provide a visual indication of when all of ethanol
has distilled over to the receiving vessel, leaving only water in the
distillation flask.
Figure 6-2. The distillation apparatus I used for the lab session in the
book
Procedure

1. If you have not already done so, put on your splash


goggles, gloves, and protective clothing.

2. Weigh the clean, dry 100 mL graduated cylinder on your


balance and record the mass on Line A in Table 6-2.

3. Transfer 100.0 mL of ethanol to the graduated cylinder,


using the disposable pipette to add the last few drops to
bring up the volume to 100.0. Reweigh the graduated
cylinder and record the mass on Line B in Table 6-2.

4. Subtract Line A from Line B and record the result (the


mass of the 100 mL of ethanol) on Line C. Divide Line C by
100 to give the density of the ethanol in g/mL and record this
value on Line D.

5. Transfer the 100 mL of ethanol from the graduated


cylinder into the 250 mL Erlenmeyer flask and set the flask
aside. Dry the inside of the graduated cylinder thoroughly.

6. Place a drop of glycerol or other lubricant on the bulb end


of the thermometer and carefully slide the thermometer into
one hole of the two-hole stopper. Wear gloves to protect your
hands if the thermometer breaks, and press gently with a
twisting motion. Do not force it. You want the bulb end of the
thermometer to protrude a few cm below the bottom of the
stopper, making sure the range from about 70 °C to 100 °C is
visible above the stopper.

7. Lubricate one of the 75 mm glass tubes and insert it into


the other hole in the rubber stopper as described above. (If
the ends of the glass tube are not already fire-polished, fire-
polish both ends and allow the tube to cool before
proceeding.) Insert the glass tube until just a few mm
protrude below the bottom of the stopper, leaving as much
as possible of the tube protruding from the top of the
stopper, where you will connect the flexible tubing.

8. Connect one end of the flexible tubing to the glass tube


protruding from the stopper, and insert the other (fire-
polished) glass tube 1 cm or so into the other end of the
flexible tubing. (You can lubricate the glass tubes slightly
with glycerol where they connect to the flexible tubing, but
that’s generally not necessary for a temporary setup. If you
leave the glass tubes inserted into the flexible tubing, they
may eventually bind to and split the flexible tubing.)

9. Insert the rubber stopper assembly carefully into the neck


of the 250 mL flask. Press down gently to ensure a seal, but
don’t force the stopper too far into the neck of the flask.

10. Place the 100 mL graduated cylinder in the coffee can or


other container and fill the container with ice water. You
needn’t use a lot of ice, but make sure that there are still at
least a few chunks of ice floating in the water after it has
cooled down completely. You want the water level to be as
high as possible against the outside of the graduated
cylinder, but not so high that there’s any risk of any water
being transferred into the graduated cylinder.

11. Place the 250 mL Erlenmeyer flask on the hotplate,


position the coffee can as necessary, and place the free end
of the glass tube on the flexible tubing inside the graduated
cylinder, making sure that the tip is well below the level of
the ice water but not so far into the graduated cylinder that
it will be submerged by distillate as it forms.

12. Turn on the hotplate to its lowest setting and observe the
liquid in the flask. As the liquid begins to boil gently, the
thermometer reading increases to indicate the temperature
of the vapor that fills the flask. That temperature is the
boiling point of the fraction that is being vaporized, which is
an ethanol/water mixture with an ethanol concentration
higher than 70% (and approaching the 95% concentration of
the low-boiling azeotrope).

13. Continue boiling the contents of the flask gently,


adjusting the heat as necessary, to maintain the liquid at a
gentle boil. As the more concentrated ethanol boils off, the
vapor moves through the flexible tubing and into the
graduated cylinder, where it cools and condenses into liquid.

14. Observe the thermometer carefully. As long as any


ethanol remains in the distillation vessel, the temperature of
the vapor remains at the boiling point of the ethanol/water
azeotrope. When all of the ethanol has boiled off, you’ll see a
sudden increase in temperature, which indicates that you
are now boiling off pure water as steam. At that point,
immediately turn off the hotplate and carefully remove the
end of the (hot) tubing from the graduated cylinder. Place
the tubing end aside in a beaker or other safe location, as
some steam will continue to come over until the distillation
vessel has cooled.

15. Remove the graduated cylinder from the cold water bath,
being careful not to spill any of the contents or to allow any
liquid from the water bath to enter the graduated cylinder.
Dry the exterior of the graduated cylinder completely, and
place it aside to allow it to equilibrate to room temperature.

16. After the graduated cylinder has reached room


temperature, make sure the exterior is completely dry and
weigh the graduated cylinder with its contents. Record the
combined mass on Line E of Table 6-2.

17. Subtract the empty mass of the graduated cylinder (Line


A) to determine the mass of the distillate. Enter that mass
on Line F of Table 6-1.

18. Determine the volume of distillate in the graduated


cylinder and record that value as accurately as possible on
Line G of Table 6-1.

19. Determine the density of the distillate by dividing its mass


by its volume. Enter the value you calculate on Line H of
Table 6-1.

20. From the density you calculated in the previous step,


determine the concentration of ethanol in the distillate by
looking up density values for aqueous ethanol solutions in a
handbook or other source. (One convenient source is
the Wikipedia Ethanol data page.)

Table 6-2. Distillation of ethanol – observed and calculated data


Item Data Disposal
A. mass of 100 mL graduated cylinder ___.___ g You can retain
the distilled
B. mass of 100 mL graduated cylinder + 100.0 mL 70% ethanol ___.___ g
ethanol and
C. mass of 100.0 mL 70% ethanol (B – A) ___.___ g use it as fuel
for your
D. density of 70% ethanol (C / 100) ___.___ g/mL
alcohol lamp
E. mass of graduated cylinder with distillate ___.___ g or for any
F. mass of distillate (E – A) ___.___ g other purpose
that requires
G. volume of distillate ___.___ mL denatured
H. density of distilled ethanol (F / H) ___.___ g/mL ethanol where
concentration
is not critical, or you can safely pour it down the drain. If you retain
the distillate, label it properly.
Optional Activities
If you have time and the required materials, consider performing these
optional activities:

 Repeat the distillation, transferring the original distillate


to the distillation vessel for re-distillation.
 Run the distillation a third time, and determine the
density (and ethanol concentration) of the third distillate.

Review Questions
Q1: Would you expect the density of the distilled ethanol solution to be
lower or higher than the density of the original ethanol solution? Why?
Q2: Using values for the densities of various concentrations of ethanol
and water that you obtain from the CRC handbook or a reliable on-line
source, estimate the ethanol concentrations in the original solution
and the distillate.
Q3: A drugstore offers denatured ethanol in concentrations of 70%,
95%, and 99% by weight. The 70% and 95% solutions are relatively
inexpensive, but the 99% solution is very costly. Why?
Q4: Distilling wine produces not colorless pure ethanol as you might
expect, but brandy, which is deeply colored and contains complex
flavors. Why?
August 24, 2009

49 RESPONSES TO LABORATORY 6.2: DISTILLATION: PURIFY ETHANOL

Distillation Theory
Summary
When you heat up a mixture of liquids, the more volitile components will tend to come off first.
There is a bit of overlap (so it is never pure), but generally we can seperate the ethanol from the
water and other impurities present. The more alcohol in the liquid, the more alcohol will be in
the vapour, so multiple distillations allow us to increase the strength & purity right up to 97.2%
The concept of distillation is really quite simple.

 Mike Nixon has compiled an excellent pdf "Distillation - How it Works"


explaining all this a lot clearer than what i have below. Right down to plate
theory for heat exchangers, without making it complicated. Either e-mail
Mike for it, or download it from me here (95 Kb, 16 Nov 99). A more detailed
explanation is in his new book "The Compleat Distiller"
at http://www.amphora-society.com.

 Another explaination, geared towards distilling ethanol from agricultural


products,is Purdue University's
http://persephone.agcom.purdue.edu/AgCom/Pubs/AE/AE-117.html note
"Alcohol distillation : basic principles, equipment, performance relationships,
and safety".

 The University of Akron has a slide show covering the basics too.

 See also : "The Brewery's" Technical Library for articles on brewing related
topics (extremely comprehensive)., and
 AllTech's company homepage has much good literature.

Here's a some-what simplified explaination ....(thanks to Mike)

When you have a mixture of liquids, each with its own boiling point when pure, then
the boiling point of the mix will lie somewhere in the middle, and this will depend on
the relative concentrations of each liquid. Pure water boils at 100 °C, and pure ethanol
boils at 78.5 °C, but a mixture of water and ethanol will boil at some point in between.
The major point about distillation is that when a mixture like that boils, then the
vapour given off is richer in the most volatile component, and when that vapour
condenses then the resulting liquid has a lower boiling point than the mix it came
from. By repeating this boiling and recondensation process up a column, using
packing to hold the condensed liquid at each stage, you can separate the components
more and more.

So if you have a mixture of liquids each with a different boiling point, then you heat
the mixture, it will heat up until the new intermediate boiling point is reached. When
you first start a distilling run, the packing in the column will be at room temperature,
so vapour given off by the boiler condenses on the first cool packing it reaches. In
condensing, the vapour gives up a lot of heat, and this warms that packing until the
liquid on it boils again. However, this liquid is richer in volatiles than the mix in the
boiler, so its boiling point is lower. When it does boil again, from the heat given off by
more condensing vapour, what you get is even richer in those most volatile
components. This process of boiling and condensing continues up the column and,
because the condensed liquid is always getting richer in volatiles, the temperature
gradually falls the higher you go. The temperature at any point is governed solely by
the boiling point of that liquid mix, and any attempt to interfere with that process will
disrupt the separation that Nature is carrying out automatically.

In contrast, the boiling point of the mix left in the boiler will very slowly start to rise
as it is left with less and less of the most volatile components.

If you started with a mixture (fermented wash) that is mostly water & ethanol, with
trace amounts of methanol, propanol, etc. then the net result will be that the most
volatile components will tend to rise in greater quantity up the column than their less
volatile cousins, and will be found in greatest concentration at the top. This would
mean that methanol, the most volatile of the lot, will win the race and you will able to
collect it and set it aside. This continues until you have collected all of the "heads"
(components that are more volatile than ethanol), and you can then collect just ethanol
with a trace of water. You cannot get rid of that small amount of water, as once you
reach a mix of 97.2% ethanol/water, with a boiling point of 78.2 °C, then you have
reached a stable mix that no amount of re-boiling and re-condensation can change (at
normal atmospheric pressure).

Once you have collected the main bulk of ethanol, then the components that are less
volatile than ethanol, such as propanol and the bigger organic molecules, will start to
reach the top, and you will have arrived at the stage called the "tails". These "tails"
may be recycled in the next batch you do, for they still contain a lot of ethanol, or a
proportion may be retained as they contain many of the compounds that give a spirit a
distinctive flavour, like whiskey or rum.

Note that you are not changing any part of your original brew - you're not "making"
the alcohol, or converting it to something else or nasty. All you are doing is
concentrating off the original brew into its various parts. There is no more methanol
after you finish than what you started with. What does happen though, is that because
most of the methanol comes off at once (first up), it is highly concentrated, and can
damage you. You definately don't want to be sampling the first portion of distillate
that you collect. But once you have thrown away this part, you have guaranteed that
the remaining distillate is safe enough to partake of.

http://homedistiller.org This page last modified Thu, 03 Aug 2017 22:10:49 -0700
Distillation
From Wikipedia, the free encyclopedia
Jump to navigationJump to search
For other uses, see Distillation (disambiguation).
"Distiller" and "Distillery" redirect here. For other uses, see Distiller (disambiguation) and Distillery
(disambiguation).
"Distill" redirects here. For the album by Bill Laswell, see Distill (album).

Laboratory display of distillation: 1: A source of heat 2: Still pot 3: Still head 4: Thermometer/Boiling point
temperature 5: Condenser 6: Cooling water in 7: Cooling water out 8: Distillate/receiving flask 9: Vacuum/gas
inlet 10: Still receiver 11: Heat control 12: Stirrer speed control 13: Stirrer/heat plate 14: Heating (Oil/sand)
bath 15: Stirring means e.g. (shown), boiling chips or mechanical stirrer 16: Cooling bath.[1]

Distillation is the process of separating the components or substances from a liquid mixture by
using selective boiling and condensation. Distillation may result in essentially complete separation
(nearly pure components), or it may be a partial separation that increases the concentration of
selected components in the mixture. In either case, the process exploits differences in the volatility of
the mixture's components. In industrial chemistry, distillation is a unit operation of practically
universal importance, but it is a physical separation process, not a chemical reaction.
Distillation has many applications. For example:

 Distillation of fermented products produces distilled beverages with a high alcohol content or
separates out other fermentation products of commercial value.

 Distillation is an effective and traditional method of desalination.

 In the fossil fuel industry, oil stabilization is a form of partial distillation that reduces vapor
pressure of crude oil, thereby making it safe for storage and transport as well as reducing the
atmospheric emissions of volatile hydrocarbons. In midstream operations at oil refineries,
distillation is a major class of operation for transforming crude oil into fuels and chemical feed
stocks.

 Cryogenic distillation leads to the separation of air into its components –


notably oxygen, nitrogen, and argon – for industrial use.

 In the field of industrial chemistry, large amounts of crude liquid products of chemical
synthesis are distilled to separate them, either from other products, from impurities, or from
unreacted starting materials.
An installation used for distillation, especially of distilled beverages, is called a distillery. The
distillation equipment at a distillery is a still.

Contents

 1History

 2Applications of distillation

 3Idealized distillation model

o 3.1Batch or differential distillation

o 3.2Continuous distillation

o 3.3General improvements

 4Laboratory scale distillation

o 4.1Simple distillation

o 4.2Fractional distillation

o 4.3Steam distillation

o 4.4Vacuum distillation

o 4.5Air-sensitive vacuum distillation

o 4.6Short path distillation

o 4.7Zone distillation

o 4.8Other types

 5Azeotropic distillation

o 5.1Breaking an azeotrope with unidirectional pressure manipulation

o 5.2Pressure-swing distillation

 6Industrial distillation

o 6.1Multi-effect distillation

 7Distillation in food processing

o 7.1Distilled beverages
 8Gallery

 9See also

 10References

 11Cited sources

 12Further reading

 13External links

History[edit]
See also: Distilled beverage

Distillation equipment used by the 3rd century alchemist Zosimos of Panopolis,[2][3] from the Byzantine
Greek manuscript Parisinus graces.[4]

Early evidence of distillation was found on Akkadian tablets dated circa 1200 BC describing
perfumery operations. The tablets provided textual evidence that an early primitive form of distillation
was known to the Babylonians of ancient Mesopotamia.[5] Early evidence of distillation was also
found related to alchemists working in Alexandria in Roman Egypt in the 1st century.[6] Distilled
water has been in use since at least c. 200, when Alexander of Aphrodisias described the process.[7]
[8]
Work on distilling other liquids continued in early Byzantine Egypt under Zosimus of Panopolis in
the 3rd century. Distillation was practiced in the ancient Indian subcontinent, which is evident from
baked clay retorts and receivers found at Taxila and Charsadda in modern Pakistan, dating back to
the early centuries of the Common Era. These "Gandhara stills" were only capable of producing very
weak liquor, as there was no efficient means of collecting the vapors at low heat.[9] Distillation in
China may have begun during the Eastern Han dynasty (1st–2nd centuries), but the distillation of
beverages began in the Jin (12th–13th centuries) and Southern Song (10th–13th centuries)
dynasties, according to archaeological evidence. [10]
Clear evidence of the distillation of alcohol comes from the Arab chemist Al-Kindi in 9th-century Iraq.
[11][12][13]
The process later spread to Italy,[11][9] where it was described by the School of Salerno in the
12th century.[6][14] Fractional distillation was developed by Tadeo Alderotti in the 13th century.[15] A still
was found in an archaeological site in Qinglong, Hebei province, in China, dating back to the 12th
century. Distilled beverages were common during the Yuan dynasty (13th–14th centuries).[10]
In 1500, German alchemist Hieronymus Braunschweig published Liber de arte destillandi (The Book
of the Art of Distillation),[16] the first book solely dedicated to the subject of distillation, followed in
1512 by a much expanded version. In 1651, John French published The Art of Distillation,[17] the first
major English compendium on the practice, but it has been claimed [18] that much of it derives from
Braunschweig's work. This includes diagrams with people in them showing the industrial rather than
bench scale of the operation.
Hieronymus Brunschwig's Liber de arte Distillandi de Compositis(Strassburg, 1512) Science History Institute

A retort

Distillation

Old Ukrainian vodka still

Simple liqueur distillation in East Timor

As alchemy evolved into the science of chemistry, vessels called retorts became used for
distillations. Both alembics and retorts are forms of glassware with long necks pointing to the side at
a downward angle to act as air-cooled condensers to condense the distillate and let it drip downward
for collection. Later, copper alembics were invented. Riveted joints were often kept tight by using
various mixtures, for instance a dough made of rye flour. [19] These alembics often featured a cooling
system around the beak, using cold water, for instance, which made the condensation of alcohol
more efficient. These were called pot stills. Today, the retorts and pot stills have been largely
supplanted by more efficient distillation methods in most industrial processes. However, the pot still
is still widely used for the elaboration of some fine alcohols, such as cognac, Scotch whisky, Irish
whiskey, tequila, and some vodkas. Pot stills made of various materials (wood, clay, stainless steel)
are also used by bootleggers in various countries. Small pot stills are also sold for use in the
domestic production[20] of flower water or essential oils.
Early forms of distillation involved batch processes using one vaporization and one condensation.
Purity was improved by further distillation of the condensate. Greater volumes were processed by
simply repeating the distillation. Chemists reportedly carried out as many as 500 to 600 distillations
in order to obtain a pure compound.[21]
In the early 19th century, the basics of modern techniques, including pre-heating and reflux, were
developed.[21] In 1822, Anthony Perrier developed one of the first continuous stills, and then, in 1826,
Robert Stein improved that design to make his patent still. In 1830, Aeneas Coffey got a patent for
improving the design even further.[22] Coffey's continuous still may be regarded as the archetype of
modern petrochemical units. The French engineer Armand Savalle developed his steam regulator
around 1846.[23] In 1877, Ernest Solvay was granted a U.S. Patent for a tray column
for ammonia distillation,[24] and the same and subsequent years saw developments in this theme for
oils and spirits.
With the emergence of chemical engineering as a discipline at the end of the 19th century, scientific
rather than empirical methods could be applied. The developing petroleum industry in the early 20th
century provided the impetus for the development of accurate design methods, such as
the McCabe–Thiele method by Ernest Thiele and the Fenske equation. The availability of powerful
computers also allowed direct computer simulations of distillation columns.

Applications of distillation[edit]
The application of distillation can roughly be divided into four groups: laboratory scale, industrial
distillation, distillation of herbs for perfumery and medicinals (herbal distillate), and food processing.
The latter two are distinctively different from the former two in that distillation is not used as a true
purification method but more to transfer all volatiles from the source materials to the distillate in the
processing of beverages and herbs.
The main difference between laboratory scale distillation and industrial distillation is that laboratory
scale distillation is often performed on a batch basis, whereas industrial distillation often occurs
continuously. In batch distillation, the composition of the source material, the vapors of the distilling
compounds, and the distillate change during the distillation. In batch distillation, a still is charged
(supplied) with a batch of feed mixture, which is then separated into its component fractions, which
are collected sequentially from most volatile to less volatile, with the bottoms – remaining least or
non-volatile fraction – removed at the end. The still can then be recharged and the process
repeated.
In continuous distillation, the source materials, vapors, and distillate are kept at a constant
composition by carefully replenishing the source material and removing fractions from both vapor
and liquid in the system. This results in a more detailed control of the separation process.

Idealized distillation model[edit]


The boiling point of a liquid is the temperature at which the vapor pressure of the liquid equals the
pressure around the liquid, enabling bubbles to form without being crushed. A special case is
the normal boiling point, where the vapor pressure of the liquid equals the ambient atmospheric
pressure.
It is a common misconception that in a liquid mixture at a given pressure, each component boils at
the boiling point corresponding to the given pressure, allowing the vapors of each component to
collect separately and purely. However, this does not occur, even in an idealized system. Idealized
models of distillation are essentially governed by Raoult's law and Dalton's law and assume
that vapor–liquid equilibria are attained.
Raoult's law states that the vapor pressure of a solution is dependent on 1) the vapor pressure of
each chemical component in the solution and 2) the fraction of solution each component makes up,
a.k.a. the mole fraction. This law applies to ideal solutions, or solutions that have different
components but whose molecular interactions are the same as or very similar to pure solutions.
Dalton's law states that the total pressure is the sum of the partial pressures of each individual
component in the mixture. When a multi-component liquid is heated, the vapor pressure of each
component will rise, thus causing the total vapor pressure to rise. When the total vapor pressure
reaches the pressure surrounding the liquid, boiling occurs and liquid turns to gas throughout the
bulk of the liquid. Note that a mixture with a given composition has one boiling point at a given
pressure when the components are mutually soluble. A mixture of constant composition does not
have multiple boiling points.
An implication of one boiling point is that lighter components never cleanly "boil first". At boiling point,
all volatile components boil, but for a component, its percentage in the vapor is the same as its
percentage of the total vapor pressure. Lighter components have a higher partial pressure and, thus,
are concentrated in the vapor, but heavier volatile components also have a (smaller) partial pressure
and necessarily vaporize also, albeit at a lower concentration in the vapor. Indeed, batch distillation
and fractionation succeed by varying the composition of the mixture. In batch distillation, the batch
vaporizes, which changes its composition; in fractionation, liquid higher in the fractionation column
contains more lights and boils at lower temperatures. Therefore, starting from a given mixture, it
appears to have a boiling range instead of a boiling point, although this is because its composition
changes: each intermediate mixture has its own, singular boiling point.
The idealized model is accurate in the case of chemically similar liquids, such
as benzene and toluene. In other cases, severe deviations from Raoult's law and Dalton's law are
observed, most famously in the mixture of ethanol and water. These compounds, when heated
together, form an azeotrope, which is when the vapor phase and liquid phase contain the same
composition. Although there are computational methods that can be used to estimate the behavior of
a mixture of arbitrary components, the only way to obtain accurate vapor–liquid equilibrium data is
by measurement.
It is not possible to completely purify a mixture of components by distillation, as this would require
each component in the mixture to have a zero partial pressure. If ultra-pure products are the goal,
then further chemical separation must be applied. When a binary mixture is vaporized and the other
component, e.g., a salt, has zero partial pressure for practical purposes, the process is simpler.
Batch or differential distillation[edit]

A batch still showing the separation of A and B.

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 that 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 in 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 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 of 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 distillation[edit]
Main article: Continuous distillation
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 occur over time during the operation. Continuous distillation produces a minimum of two
output fractions, including at least one volatile distillate fraction, which has boiled and been
separately captured as a vapor and then 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.
For any source material of specific composition, the main variables that affect the purity of products
in continuous distillation are the reflux ratio and the number of theoretical equilibrium stages, in
practice determined by 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.
General improvements[edit]
Both batch and continuous distillations can be improved by making use of a fractionating column on
top of the distillation flask. The column improves separation by providing a larger surface area for the
vapor and condensate to come into contact. This helps it remain at equilibrium for as long as
possible. The column can even consist of small subsystems ('trays' or 'dishes') which all contain an
enriched, boiling liquid mixture, all with their own vapor–liquid equilibrium.
There are differences between laboratory-scale and industrial-scale fractionating columns, but the
principles are the same. Examples of laboratory-scale fractionating columns (in increasing efficiency)
include

 Air condenser

 Vigreux column (usually laboratory scale only)

 Packed column (packed with glass beads, metal pieces, or other chemically inert material)

 Spinning band distillation system.

Laboratory scale distillation[edit]

Typical laboratory fractional distillation unit

Laboratory scale distillations are almost exclusively run as batch distillations. The device used in
distillation, sometimes referred to as a still, consists at a minimum of a reboiler or pot in which the
source material is heated, a condenser in which the heated vapour is cooled back to the
liquid state, and a receiver in which the concentrated or purified liquid, called the distillate, is
collected. Several laboratory scale techniques for distillation exist (see also distillation types).
Simple distillation[edit]
In simple distillation, the vapor is immediately channeled into a condenser. Consequently, the
distillate is not pure but rather its composition is identical to the composition of the vapors at the
given temperature and pressure. That concentration follows Raoult's law.
As a result, simple distillation is effective only when the liquid boiling points differ greatly (rule of
thumb is 25 °C)[25] or when separating liquids from non-volatile solids or oils. For these cases, the
vapor pressures of the components are usually different enough that the distillate may be sufficiently
pure for its intended purpose.
Fractional distillation[edit]
Main article: Fractional distillation
For many cases, the boiling points of the components in the mixture will be sufficiently close
that Raoult's law must be taken into consideration. Therefore, fractional distillation must be used in
order to separate the components by repeated vaporization-condensation cycles within a packed
fractionating column. This separation, by successive distillations, is also referred to as rectification.
[26]

As the solution to be purified is heated, its vapors rise to the fractionating column. As it rises, it cools,
condensing on the condenser walls and the surfaces of the packing material. Here, the condensate
continues to be heated by the rising hot vapors; it vaporizes once more. However, the composition of
the fresh vapors are determined once again by Raoult's law. Each vaporization-condensation cycle
(called a theoretical plate) will yield a purer solution of the more volatile component. [27] In reality, each
cycle at a given temperature does not occur at exactly the same position in the fractionating
column; theoretical plate is thus a concept rather than an accurate description.
More theoretical plates lead to better separations. A spinning band distillation system uses a
spinning band of Teflon or metal to force the rising vapors into close contact with the descending
condensate, increasing the number of theoretical plates.[28]
Steam distillation[edit]
Main article: Steam distillation
Like vacuum distillation, steam distillation is a method for distilling compounds which are heat-
sensitive.[29] The temperature of the steam is easier to control than the surface of a heating element,
and allows a high rate of heat transfer without heating at a very high temperature. This process
involves bubbling steam through a heated mixture of the raw material. By Raoult's law, some of the
target compound will vaporize (in accordance with its partial pressure). The vapor mixture is cooled
and condensed, usually yielding a layer of oil and a layer of water.
Steam distillation of various aromatic herbs and flowers can result in two products; an essential
oil as well as a watery herbal distillate. The essential oils are often used in perfumery
and aromatherapy while the watery distillates have many applications in aromatherapy, food
processing and skin care.

Dimethyl sulfoxide usually boils at 189 °C. Under a vacuum, it distills off into the receiver at only 70 °C.
Perkin triangle distillation setup
1: Stirrer bar/anti-bumping granules 2:Still pot 3: Fractionating column 4:Thermometer/Boiling point
temperature 5: Teflon tap 1 6: Cold finger 7: Cooling water out 8: Cooling water in 9: Teflon tap
2 10: Vacuum/gas inlet 11: Teflon tap 3 12: Still receiver

Vacuum distillation[edit]
Main article: Vacuum distillation
Some compounds have very high boiling points. To boil such compounds, it is often better to lower
the pressure at which such compounds are boiled instead of increasing the temperature. Once the
pressure is lowered to the vapor pressure of the compound (at the given temperature), boiling and
the rest of the distillation process can commence. This technique is referred to as vacuum
distillation and it is commonly found in the laboratory in the form of the rotary evaporator.
This technique is also very useful for compounds which boil beyond their decomposition
temperatureat atmospheric pressure and which would therefore be decomposed by any attempt to
boil them under atmospheric pressure.
Molecular distillation is vacuum distillation below the pressure of 0.01 torr. 0.01 torr is one order of
magnitude above high vacuum, where fluids are in the free molecular flow regime, i.e. the mean free
path of molecules is comparable to the size of the equipment. The gaseous phase no longer exerts
significant pressure on the substance to be evaporated, and consequently, rate of evaporation no
longer depends on pressure. That is, because the continuum assumptions of fluid dynamics no
longer apply, mass transport is governed by molecular dynamics rather than fluid dynamics. Thus, a
short path between the hot surface and the cold surface is necessary, typically by suspending a hot
plate covered with a film of feed next to a cold plate with a line of sight in between. Molecular
distillation is used industrially for purification of oils.
Air-sensitive vacuum distillation[edit]
Some compounds have high boiling points as well as being air sensitive. A simple vacuum distillation
system as exemplified above can be used, whereby the vacuum is replaced with an inert gas after
the distillation is complete. However, this is a less satisfactory system if one desires to collect
fractions under a reduced pressure. To do this a "cow" or "pig" adaptor can be added to the end of
the condenser, or for better results or for very air sensitive compounds a Perkin triangle apparatus
can be used.
The Perkin triangle, has means via a series of glass or Teflon taps to allows fractions to be isolated
from the rest of the still, without the main body of the distillation being removed from either the
vacuum or heat source, and thus can remain in a state of reflux. To do this, the sample is first
isolated from the vacuum by means of the taps, the vacuum over the sample is then replaced with
an inert gas (such as nitrogen or argon) and can then be stoppered and removed. A fresh collection
vessel can then be added to the system, evacuated and linked back into the distillation system via
the taps to collect a second fraction, and so on, until all fractions have been collected.
Short path distillation[edit]
Short path vacuum distillation apparatus with vertical condenser (cold finger), to minimize the distillation
path; 1: Still pot with stirrer bar/anti-bumping granules 2: Cold finger – bent to direct condensate 3: Cooling
water out 4:cooling water in 5: Vacuum/gas inlet 6:Distillate flask/distillate.

Short path distillation is a distillation technique that involves the distillate travelling a short distance,
often only a few centimeters, and is normally done at reduced pressure. [30] A classic example would
be a distillation involving the distillate travelling from one glass bulb to another, without the need for a
condenser separating the two chambers. This technique is often used for compounds which are
unstable at high temperatures or to purify small amounts of compound. The advantage is that the
heating temperature can be considerably lower (at reduced pressure) than the boiling point of the
liquid at standard pressure, and the distillate only has to travel a short distance before condensing. A
short path ensures that little compound is lost on the sides of the apparatus. The Kugelrohr is a kind
of a short path distillation apparatus which often contain multiple chambers to collect distillate
fractions.
Zone distillation[edit]
Zone distillation is a distillation process in long container with partial melting of refined matter in
moving liquid zone and condensation of vapor in the solid phase at condensate pulling in cold area.
The process is worked in theory. When zone heater is moving from the top to the bottom of the
container then solid condensate with irregular impurity distribution is forming. Then most pure part of
the condensate may be extracted as product. The process may be iterated many times by moving
(without turnover) the received condensate to the bottom part of the container on the place of refined
matter. The irregular impurity distribution in the condensate (that is efficiency of purification)
increases with number of repetitions of the process. Zone distillation is a distillation analog of zone
recrystallization. Impurity distribution in the condensate is described by known equations of zone
recrystallization with various numbers of iteration of process – with replacement distribution efficient
k of crystallization on separation factor α of distillation. [31][32][33]
Other types[edit]

 The process of reactive distillation involves using the reaction vessel as the still. In this
process, the product is usually significantly lower-boiling than its reactants. As the product is
formed from the reactants, it is vaporized and removed from the reaction mixture. This technique
is an example of a continuous vs. a batch process; advantages include less downtime to charge
the reaction vessel with starting material, and less workup. Distillation "over a reactant" could be
classified as a reactive distillation. It is typically used to remove volatile impurity from the
distallation feed. For example, a little lime may be added to remove carbon dioxide from water
followed by a second distillation with a little sulfuric acid added to remove traces of ammonia.

 Catalytic distillation is the process by which the reactants are catalyzed while being distilled
to continuously separate the products from the reactants. This method is used to assist
equilibrium reactions reach completion.

 Pervaporation is a method for the separation of mixtures of liquids by partial vaporization


through a non-porous membrane.

 Extractive distillation is defined as distillation in the presence of a miscible, high boiling,


relatively non-volatile component, the solvent, that forms no azeotrope with the other
components in the mixture.
 Flash evaporation (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,
being equivalent to a distillation with only one equilibrium stage.

 Codistillation is distillation which is performed on mixtures in which the two compounds are
not miscible. In the laboratory, the Dean-Stark apparatus is used for this purpose to remove
water from synthesis products. The Bleidner is another example with two refluximg solvents.

 Membrane distillation is a type of distillation in which vapors of a mixture to be separated are


passed through a membrane, which selectively permeates one component of mixture. Vapor
pressure difference is the driving force. It has potential applications in seawater desalination and
in removal of organic and inorganic components.
The unit process of evaporation may also be called "distillation":

 In rotary evaporation a vacuum distillation apparatus is used to remove bulk solvents from a
sample. Typically the vacuum is generated by a water aspirator or a membrane pump.

 In a kugelrohr a short path distillation apparatus is typically used (generally in combination


with a (high) vacuum) to distill high boiling (> 300 °C) compounds. The apparatus consists of an
oven in which the compound to be distilled is placed, a receiving portion which is outside of the
oven, and a means of rotating the sample. The vacuum is normally generated by using a high
vacuum pump.
Other uses:

 Dry distillation or destructive distillation, despite the name, is not truly distillation, but rather
a chemical reaction known as pyrolysis in which solid substances are heated in an inert
or reducing atmosphere and any volatile fractions, containing high-boiling liquids and products of
pyrolysis, are collected. The destructive distillation of wood to give methanol is the root of its
common name – wood alcohol.

 Freeze distillation is an analogous method of purification using freezing instead of


evaporation. It is not truly distillation, but a recrystallization where the product is the mother
liquor, and does not produce products equivalent to distillation. This process is used in the
production of ice beer and ice wine to increase ethanol and sugar content, respectively. It is also
used to produce applejack. Unlike distillation, freeze distillation concentrates poisonous
congeners rather than removing them; As a result, many countries prohibit such applejack as a
health measure. However, reducing methanol with the absorption of 4A molecular sieve is a
practical method for production.[34] Also, distillation by evaporation can separate these since they
have different boiling points.

Azeotropic distillation[edit]
Main article: Azeotropic distillation
Interactions between the components of the solution create properties unique to the solution, as
most processes entail nonideal mixtures, where Raoult's law does not hold. Such interactions can
result in a constant-boiling azeotrope which behaves as if it were a pure compound (i.e., boils at a
single temperature instead of a range). At an azeotrope, the solution contains the given component
in the same proportion as the vapor, so that evaporation does not change the purity, and distillation
does not effect separation. For example, ethyl alcohol and water form an azeotrope of 95.6% at
78.1 °C.
If the azeotrope is not considered sufficiently pure for use, there exist some techniques to break the
azeotrope to give a pure distillate. This set of techniques are known as azeotropic distillation.
Some techniques achieve this by "jumping" over the azeotropic composition (by adding another
component to create a new azeotrope, or by varying the pressure). Others work by chemically or
physically removing or sequestering the impurity. For example, to purify ethanol beyond 95%, a
drying agent (or desiccant, such as potassium carbonate) can be added to convert the soluble water
into insoluble water of crystallization. Molecular sieves are often used for this purpose as well.
Immiscible liquids, such as water and toluene, easily form azeotropes. Commonly, these azeotropes
are referred to as a low boiling azeotrope because the boiling point of the azeotrope is lower than
the boiling point of either pure component. The temperature and composition of the azeotrope is
easily predicted from the vapor pressure of the pure components, without use of Raoult's law. The
azeotrope is easily broken in a distillation set-up by using a liquid–liquid separator (a decanter) to
separate the two liquid layers that are condensed overhead. Only one of the two liquid layers is
refluxed to the distillation set-up.
High boiling azeotropes, such as a 20 weight percent mixture of hydrochloric acid in water, also
exist. As implied by the name, the boiling point of the azeotrope is greater than the boiling point of
either pure component.
To break azeotropic distillations and cross distillation boundaries, such as in the DeRosier Problem,
it is necessary to increase the composition of the light key in the distillate.
Breaking an azeotrope with unidirectional pressure manipulation [edit]
The boiling points of components in an azeotrope overlap to form a band. By exposing an azeotrope
to a vacuum or positive pressure, it's possible to bias the boiling point of one component away from
the other by exploiting the differing vapour pressure curves of each; the curves may overlap at the
azeotropic point, but are unlikely to be remain identical further along the pressure axis either side of
the azeotropic point. When the bias is great enough, the two boiling points no longer overlap and so
the azeotropic band disappears.
This method can remove the need to add other chemicals to a distillation, but it has two potential
drawbacks.
Under negative pressure, power for a vacuum source is needed and the reduced boiling points of
the distillates requires that the condenser be run cooler to prevent distillate vapours being lost to the
vacuum source. Increased cooling demands will often require additional energy and possibly new
equipment or a change of coolant.
Alternatively, if positive pressures are required, standard glassware can not be used, energy must be
used for pressurization and there is a higher chance of side reactions occurring in the distillation,
such as decomposition, due to the higher temperatures required to effect boiling.
A unidirectional distillation will rely on a pressure change in one direction, either positive or negative.
Pressure-swing distillation[edit]
Further information: Azeotrope § Pressure swing distillation
Pressure-swing distillation is essentially the same as the unidirectional distillation used to break
azeotropic mixtures, but here both positive and negative pressures may be employed.
This improves the selectivity of the distillation and allows a chemist to optimize distillation by
avoiding extremes of pressure and temperature that waste energy. This is particularly important in
commercial applications.
One example of the application of pressure-swing distillation is during the industrial purification
of ethyl acetate after its catalytic synthesis from ethanol.

Industrial distillation[edit]

Typical industrial distillation towers

Main article: Continuous distillation


Large scale industrial distillation applications include both batch and continuous fractional,
vacuum, azeotropic, extractive, and steam distillation. The most widely used industrial applications of
continuous, steady-state fractional distillation are in petroleum
refineries, petrochemical and chemical plants and natural gas processing plants.
To control and optimize such industrial distillation, a standardized laboratory method, ASTM D86, is
established. This test method extends to the atmospheric distillation of petroleum products using a
laboratory batch distillation unit to quantitatively determine the boiling range characteristics of
petroleum products.
Industrial distillation[26][35] is typically performed in large, vertical cylindrical columns known
as distillation towers or distillation columnswith diameters ranging from about 65 centimeters to
16 meters and heights ranging from about 6 meters to 90 meters or more. When the process feed
has a diverse composition, as in distilling crude oil, liquid outlets at intervals up the column allow for
the withdrawal of different fractions or products having different boiling points or boiling ranges. The
"lightest" products (those with the lowest boiling point) exit from the top of the columns and the
"heaviest" products (those with the highest boiling point) exit from the bottom of the column and are
often called the bottoms.

Diagram of a typical industrial distillation tower

Industrial towers use reflux to achieve a more complete separation of products. Reflux refers to the
portion of the condensed overhead liquid product from a distillation or fractionation tower that is
returned to the upper part of the tower as shown in the schematic diagram of a typical, large-scale
industrial distillation tower. Inside the tower, the downflowing reflux liquid provides cooling and
condensation of the upflowing vapors thereby increasing the efficiency of the distillation tower. The
more reflux that is provided for a given number of theoretical plates, the better the tower's separation
of lower boiling materials from higher boiling materials. Alternatively, the more reflux that is provided
for a given desired separation, the fewer the number of theoretical plates required. Chemical
engineers must choose what combination of reflux rate and number of plates is both economically
and physically feasible for the products purified in the distillation column.
Such industrial fractionating towers are also used in cryogenic air separation, producing liquid
oxygen, liquid nitrogen, and high purity argon. Distillation of chlorosilanes also enables the
production of high-purity silicon for use as a semiconductor.
Section of an industrial distillation tower showing detail of trays with bubble caps

Design and operation of a distillation tower depends on the feed and desired products. Given a
simple, binary component feed, analytical methods such as the McCabe–Thiele method[26][36] or
the Fenske equation[26] can be used. For a multi-component feed, simulation models are used both
for design and operation. Moreover, the efficiencies of the vapor–liquid contact devices (referred to
as "plates" or "trays") used in distillation towers are typically lower than that of a theoretical 100%
efficient equilibrium stage. Hence, a distillation tower needs more trays than the number of
theoretical vapor–liquid equilibrium stages. A variety of models have been postulated to estimate tray
efficiencies.
In modern industrial uses, a packing material is used in the column instead of trays when low
pressure drops across the column are required. Other factors that favor packing are: vacuum
systems, smaller diameter columns, corrosive systems, systems prone to foaming, systems
requiring low liquid holdup, and batch distillation. Conversely, factors that favor plate columns are:
presence of solids in feed, high liquid rates, large column diameters, complex columns, columns with
wide feed composition variation, columns with a chemical reaction, absorption columns, columns
limited by foundation weight tolerance, low liquid rate, large turn-down ratio and those processes
subject to process surges.

Large-scale, industrial vacuum distillation column[37]

This packing material 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 pass
across this wetted surface, where mass transfertakes place. Unlike conventional tray distillation in
which every tray represents a separate point of vapor–liquid equilibrium, the vapor–liquid equilibrium
curve in a packed column is continuous. However, when modeling packed columns, it is useful to
compute a number of "theoretical stages" to denote the separation efficiency of the packed column
with respect to more traditional trays. Differently shaped packings have different surface areas and
void space between packings. Both of these factors affect packing performance.
Another factor in addition to the packing shape and surface area that affects the performance of
random or structured packing is the liquid and vapor distribution entering the packed bed. The
number of theoretical stages required to make a given separation is calculated using a specific vapor
to liquid ratio. If the liquid and vapor are not evenly distributed across the superficial tower area as it
enters the packed bed, the liquid to vapor ratio will not be correct in the packed bed and the required
separation will not be achieved. The packing will appear to not be working properly. The height
equivalent to a theoretical plate (HETP) will be greater than expected. The problem is not the
packing itself but the mal-distribution of the fluids entering the packed bed. Liquid mal-distribution is
more frequently the problem than vapor. The design of the liquid distributors used to introduce the
feed and reflux to a packed bed is critical to making the packing perform to it maximum efficiency.
Methods of evaluating the effectiveness of a liquid distributor to evenly distribute the liquid entering a
packed bed can be found in references.[38][39]Considerable work has been done on this topic by
Fractionation Research, Inc. (commonly known as FRI). [40]
Multi-effect distillation[edit]
The goal of multi-effect distillation is to increase the energy efficiency of the process, for use in
desalination, or in some cases one stage in the production of ultrapure water. The number of effects
is inversely proportional to the kW·h/m3 of water recovered figure, and refers to the volume of water
recovered per unit of energy compared with single-effect distillation. One effect is roughly
636 kW·h/m3.

 Multi-stage flash distillation can achieve more than 20 effects with thermal energy input, as
mentioned in the article.

 Vapor compression evaporation – Commercial large-scale units can achieve around 72


effects with electrical energy input, according to manufacturers.
There are many other types of multi-effect distillation processes, including one referred to as simply
multi-effect distillation (MED), in which multiple chambers, with intervening heat exchangers, are
employed.

Distillation in food processing[edit]


Distilled beverages[edit]
Main article: Distilled beverage
Carbohydrate-containing plant materials are allowed to ferment, producing a dilute solution of
ethanol in the process. Spirits such as whiskey and rum are prepared by distilling these dilute
solutions of ethanol. Components other than ethanol, including water, esters, and other alcohols, are
collected in the condensate, which account for the flavor of the beverage. Some of these beverages
are then stored in barrels or other containers to acquire more flavor compounds and characteristic
flavors.

Gallery[edit]

Chemistry in its beginnings used retorts as laboratory equipment exclusively for


distillation processes.

A simple set-up to distill dry and oxygen-free toluene.


Diagram of an industrial-scale vacuum distillation column as commonly used in oil
refineries

A rotary evaporator is able to distill solvents more quickly at lower temperatures


through the use of a vacuum.

Distillation using semi-microscale apparatus. The jointless design eliminates the need to
fit pieces together. The pear-shaped flask allows the last drop of residue to be removed,
compared with a similarly-sized round-bottom flask The small holdup volume prevents
losses. A pig is used to channel the various distillates into three receiving flasks. If
necessary the distillation can be carried out under vacuum using the vacuum adapter at
the pig.

See also[edit]

 Distillation portal

 Atmospheric distillation of crude oil

 Clyssus

 Fragrance extraction

 Microdistillery

 Sublimation

 Dixon rings

 Random column packing

References[edit]
1. ^ Harwood & Moody 1989, pp. 141–143

2. ^ Gildemeister, E.; Hoffman, Fr.; translated by Edward Kremers (1913). The Volatile Oils. 1.
New York: Wiley. p. 203.
3. ^ Bryan H. Bunch; Alexander Hellemans (2004). The History of Science and Technology.
Houghton Mifflin Harcourt. p. 88. ISBN 978-0-618-22123-3.

4. ^ Berthelot, Marcelin (1887–1888) Collection des anciens alchimistes grecs. 3 vol., Paris, p.
161

5. ^ Levey, Martin (1959). Chemistry and Chemical Technology in Ancient


Mesopotamia. Elsevier. p. 36. As already mentioned, the textual evidence for Sumero-Babylonian
distillation is disclosed in a group of Akkadian tablets describing perfumery operations, dated ca. 1200
B.C.

6. ^ Jump up to:a b Forbes 1970, pp. 57, 89

7. ^ Taylor, F. (1945). "The evolution of the still". Annals of Science. 5 (3):


185. doi:10.1080/00033794500201451.

8. ^ Berthelot, M. P. E. M. (1893). "The Discovery of Alcohol and Distillation". The Popular


Science Monthly. XLIII: 85–94. Archived from the original on 29 November 2017.

9. ^ Jump up to:a b Habib, Irfan (2011), Economic History of Medieval India, 1200–1500. Pearson
Education. p. 55. ISBN 9788131727911

10. ^ Jump up to:a b Haw, Stephen G. (2012). "Wine, women and poison". Marco Polo in China.
Routledge. pp. 147–148. ISBN 978-1-134-27542-7. The earliest possible period seems to be the
Eastern Han dynasty ... the most likely period for the beginning of true distillation of spirits for drinking
in China is during the Jin and Southern Song dynasties

11. ^ Jump up to:a b al-Hassan, Ahmad Y. (2001), Science and Technology in Islam: Technology
and applied sciences. UNESCO. pp. 65–69. ISBN 9789231038310

12. ^ Hassan, Ahmad Y. "Alcohol and the Distillation of Wine in Arabic Sources". History of
Science and Technology in Islam. Archived from the original on 29 December 2015. Retrieved 19
April 2014.

13. ^ The Economist: "Liquid fire – The Arabs discovered how to distil alcohol. They still do it best,
say some" Archived 22 October 2012 at the Wayback Machine. December 18, 2003

14. ^ Sarton, George (1975). Introduction to the history of science. R. E. Krieger Pub. Co.
p. 145. ISBN 978-0-88275-172-6.

15. ^ Holmyard, Eric John (1990). Alchemy. Courier Dover Publications. p. 53. ISBN 978-0-486-
26298-7.

16. ^ Braunschweig, Hieronymus (1500). Liber de arte destillandi, de Simplicibus [The Book of
the Art of Distillation] (in German).

17. ^ French, John (1651). The Art of Distillation. London: Richard Cotes.

18. ^ "Distillation". Industrial & Engineering Chemistry. 28 (6): 677.


1936. doi:10.1021/ie50318a015.

19. ^ Sealing Technique, accessed 16 November 2006.

20. ^ Traditional Alembic Pot Still Archived 21 November 2006 at the Wayback Machine.,
accessed 16 November 2006.

21. ^ Jump up to:a b Othmer, D. F. (1982) "Distillation – Some Steps in its Development", in W. F.
Furter (ed) A Century of Chemical Engineering. ISBN 0-306-40895-3

22. ^ GB 5974, Coffey, A., "Apparatus for Brewing and Distilling", published 5 August 1830, issued
5 February 1831; image Archived 4 February 2017 at the Wayback Machine.
23. ^ Forbes 1970, p. 323

24. ^ US 198699, Solvay, Ernest, "Improvement in the Ammonia-Soda Manufacture", published 2


June 1876, issued 25 December 1877

25. ^ ST07 Separation of liquid–liquid mixtures (solutions), DIDAC by IUPAC

26. ^ Jump up to:a b c d Perry, Robert H.; Green, Don W. (1984). Perry's Chemical Engineers'
Handbook(6th ed.). McGraw-Hill. ISBN 978-0-07-049479-4.

27. ^ Fractional Distillation. fandm.edu

28. ^ Spinning Band Distillation Archived 25 August 2006 at the Wayback Machine.. B/R
Instrument Corporation (accessed 8 September 2006)

29. ^ Harwood & Moody 1989, pp. 151–153

30. ^ Harwood & Moody 1989, p. 150

31. ^ Kravchenko, A. I. (2011). "Зонная дистилляция: новый метод рафинирования" [Zone


distillation: a new method of refining]. Problems of Atomic Science and Technology (in
Russian). 6 (19): 24–26.

32. ^ Kravchenko, A. I. (2014). "Zone distillation: justification". Problems of Atomic Science and
Technology. 1 (20): 64–65.

33. ^ Kravchenko, A. I. (2014). "Разработка перспективных схем зонной


дистилляции"[Design of advanced processes of zone distillation]. Perspectivnye Materialy (in
Russian) (7): 68–72.

34. ^ Study on Method of Decreasing Methanol in Apple Pomace Spirit Archived 17 May 2013 at
the Wayback Machine..

35. ^ Kister, Henry Z. (1992). Distillation Design (1st ed.). McGraw-Hill. ISBN 978-0-07-034909-4.

36. ^ Seader, J. D.; Henley, Ernest J. (1998). Separation Process Principles. New York:
Wiley. ISBN 978-0-471-58626-5.

37. ^ Energy Institute website page Archived 12 October 2007 at the Wayback Machine..
Resources.schoolscience.co.uk. Retrieved on 2014-04-20.

38. ^ Moore, F., Rukovena, F. (August 1987) Random Packing, Vapor and Liquid Distribution:
Liquid and gas distribution in commercial packed towers, Chemical Plants & Processing, Edition
Europe, pp. 11–15

39. ^ Spiegel, L (2006). "A new method to assess liquid distributor quality". Chemical Engineering
and Processing. 45 (11): 1011. doi:10.1016/j.cep.2006.05.003.

40. ^ Kunesh, John G.; Lahm, Lawrence; Yanagi, Takashi (1987). "Commercial scale experiments
that provide insight on packed tower distributors". Industrial & Engineering Chemistry
Research. 26 (9): 1845. doi:10.1021/ie00069a021.

2. Distillation

3.

4. Distillation is an important commercial process that is used in the purification


of a large variety of materials. However, before we begin a discussion of
distillation, it would probably be beneficial to define the terms that describe the
process and related properties. Many of these are terms that you are familiar
with but the exact definitions may not be known to you. Let us begin by
describing the process by which a substance is transformed from the condensed
phase to the gas phase. For a liquid, this process is called vaporization and for
a solid it is called sublimation. Both processes require heat. This is why even
on a hot day at the beach, if there is a strong breeze blowing, it may feel cool or
cold after you come out of the water. The wind facilitates the evaporation
process and you supply some of the heat that is required. All substances
regardless of whether they are liquids or solids are characterized by a vapor
pressure. The vapor pressure of a pure substance is the pressure exerted by the
substance against the external pressure which is usually atmospheric pressure.
Vapor pressure is a measure of the tendency of a condensed substance to escape
the condensed phase. The larger the vapor pressure, the greater the tendency to
escape. When the vapor pressure of a liquid substance reaches the external
pressure, the substance is observed to boil. If the external pressure is
atmospheric pressure, the temperature at which a pure substance boils is called
the normal boiling point. Solid substances are not characterized by a similar
phenomena as boiling. They simply vaporize directly into the atmosphere.
Many of you may have noticed that even on a day in which the temperature
stays below freezing, the volume of snow and ice will appear to decrease,
particularly from dark pavements on the streets. This is a consequence of the
process of sublimation. Both vaporization and sublimation are processes that
can be used to purify compounds. In order to understand how to take advantage
of these processes in purifying organic materials, we first need to learn how
pure compounds behave when they are vaporized or sublimed.

5. Let's begin by discussing the vapor pressure of a pure substance and how it
varies with temperature. Vapor pressure is an equilibrium property. If we return
to that hot windy day at the beach and consider the relative humidity in the air,
the cooling effect of the wind would be most effective if the relative humidity
was low. If the air contained a great deal of water vapor, its cooling effect
would be greatly diminished and if the relative humidity was 100%, there
would be no cooling effect. Everyone in St. Louis has experienced how long it
takes to dry off on a hot humid day. At equilibrium, the process of vaporization
is compensated by an equal amount of condensation. Incidentally, if
vaporization is an endothermic process (i.e. heat is absorbed), condensation
must be an exothermic process (i.e. heat is liberated). Now consider how vapor
pressure varies with temperature. Figure 1 illustrates that vapor pressure is a
very sensitive function of temperature. It does not increase linearly but in fact
increases exponentially with temperature. A useful "rule of thumb" is that the
vapor pressure of a substance roughly doubles for every increase in 10 °C. If
we follow the temperature dependence of vapor pressure for a substance like
water left out in an open container, we would find that the equilibrium vapor
pressure of water would increase until it reached 1 atmosphere or 101325 Pa
(101.3 kPa, 760 mm

6. Hg). At this temperature and pressure, the water would begin to boil and would
continue to do so

7. until all of the water distilled or boiled off. It is not possible to achieve a vapor
pressure greater than 1 atmosphere in a container left open to the atmosphere.
Of course, if we put a lid on the container, the vapor pressure of water or any
other substance for that matter would continue to

8.

9. Figure 1. Vapor pressure dependence on temperature for water.

10.

11. rise with temperature until the container ruptured. Elevation of the boiling point
with increase in external pressure is the principle behind the use of a pressure
cooker.

12.

13.Vacuum Distillation

14.Elevation of the boiling point with an increase in external pressure, while


important in cooking and sterilizing food or utensils, is less important in
distillation. However, it illustrates an important principle that is used in the
distillation of many materials. If the boiling point of water is increased when
the external pressure is increased, then decreasing the external pressure should
decrease the boiling point. While this is not particularly important for the
purification of water, this principle is used in the process of freeze drying, an
important commercial process. In addition, many compounds cannot be
distilled at atmospheric pressure because their boiling points are so high. At
their normal boiling points, the compounds decompose. Some of these
materials can be distilled under reduced pressure however, because the required
temperature to boil the substance can be lowered significantly. Rewording the
"rule of thumb" described above so that it is applicable here suggests that the
boiling point will be lowered by 10 °C each time the external pressure is
halved. For example, if the external pressure above a substance is reduced to
1/16 of an atmosphere by mean of a mechanical pump, the boiling point will
have been reduced four times by 10 °C for a total reduction of 40 °C (1 atm x
(1/2)(1/2)(1/2)(1/2) =1/16 atm).

15.

16.A nomograph is a useful device that can be used to estimate the boiling point of
a liquid under reduced pressure under any conditions provide either the normal
boiling point or the boiling
17.

18.(a) (b) (c)

19.

20.Figure 2. A nomograph used to estimate boiling points at reduced pressures. To


use, place a straight edge on two of the three known properties and read out the
third. Column c is in mm of mercury. An atmosphere is also equivalent to 101.3
kPa and will support a column of mercury, 76 cm (760 mm).

21.point at a some given pressure is available. To use the nomograph given the
normal boiling point, simply place a straight edge at on the temperature in the
central column of the nomograph (b). Rotating the straight edge about this
temperature will afford the expected boiling point for any number of external
pressures. Simply read the temperature and the corresponding pressure from
where the straight edge intersects the first and third columns. As an example
lets choose a normal boiling point of 400 °C. Using the nomograph in Figure 2
and this temperature for reference, rotating the straight edge about this
temperature will afford a continuous range of expected boiling points and the
required external pressures necessary to achieve the desired boiling point. At a
pressure of 6 mm, the expected boiling point would be 200 °C. Likewise, our
compound boiling at 400 °C at 1 atm would be expected to boil at 145 °C at 0.1
mm external pressure.

22.

23.Simple Distillation

24.

25.Although all of us have brought water to a boil many times, some of us may
have not realized that the temperature of pure boiling water does not change as
it distills. This is why vigorous boiling does not cook food any faster than a
slow gentle boil. The observation that the boiling point of a pure material does
not change during the course of distillation is an important property of a pure
material. The boiling point and boiling point range have been used as criteria in
confirming both the identity and purity of a substance. For example, if we
synthesized a known liquid that boiled at 120-122 °C, this value could be used
to confirm that we prepared what we were interested in and that our substance
was reasonably pure. Of course, additional criteria must also be satisfied before
the identity and purity of the liquid are known with certainty. In general, a
boiling point range of 1-2 °C is usually taken as an indication of a pure
material. You will use both of these properties later in the semester to identity
an unknown liquid.

26.

27.Occasionally, mixtures of liquids called azeotropes can be encountered that


mimic the boiling behavior of pure liquids. These mixtures when present at
specific concentrations usually distill at a constant boiling temperature and can
not be separated by distillation. Examples of such mixtures are 95% ethanol-
5% water (bp 78.1 °C), 20% acetone-80% chloroform (bp 64.7 °C), 74.1%
benzene, 7.4% water, 18.5 % ethanol (bp 64.9). The azeotropic composition
sometimes boils lower the than boiling point of its components and sometimes
higher. Mixtures of these substances at compositions other than those given
above behave as mixtures.

28.

29.Returning to our discussion of boiling water, if we were making a syrup by the


addition of sugar to boiling water, we would find that the boiling point of the
syrup would increase as the syrup begins to thicken and the sugar concentration
becomes significant. Unlike pure materials, the boiling point of an impure
liquid will change and this change is a reflection of the change in the
composition of the liquid. In fact it is this dependence of boiling point on
composition that forms the basis of using distillation for purifying liquids. We
will begin our discussion of distillation by introducing Raoult's Law, which
treats liquids in a simple and ideal, but extremely useful manner.

30.

31.
32.

33.

34.Figure 3. The apparatus used in a simple distillation. Note the position of the
thermometer bulb in the distillation head and the arrangement of the flow of the
cooling water.

35.

36.

37. = , where is the observed vapor pressure of A 1

38. = nA/(nA + nB + nC +...)

39. = vapor pressure of pure A

40.nA, nB, nC ... : number of moles of A, B, C, ...

41.

42.This relationship as defined is capable of describing the boiling point behavior


of compound A in a mixture of compounds under a variety of different
circumstances. Although this equation treats mixtures of compounds in an
oversimplified fashion and is not applicable to azeotropic compositions, it does
give a good representation of the behavior of many mixtures.
43.

44.Let's first consider a binary system (2 components) in which only one of the
two components is appreciably volatile. Our previous discussion of sugar +
water is a good example. Raoult's law states that the observed vapor pressure of
water is simply equal to the product of the mole fraction of the water present
and the vapor pressure of pure water at the temperature of the mixture. Once
the sugar-water mixture starts to boil, and continues to boil, we know that the
observed vapor pressure of the water must equal one atmosphere. Water is the
only component that is distilling. Since the mole fraction of water in a mixture
of sugar-water must be less than 1, in order for the observed vapor pressure of
water ( ) to equal one atmosphere, must be greater than one atmosphere.
This can only happen according to Figure 1 if the temperature of the mixture is
greater than 100 °C. As the distillation of water continues, the mole fraction of
the water continues to decrease thereby causing the temperature of the mixture
to increase. Remember, heat is constantly being added. If at some point the
composition of the solution becomes saturated with regards to sugar and the
sugar begins to crystallize out of solution, the composition of the solution will
become constant; removal of any additional water will simply result in the
deposit of more sugar crystals. Under these set of circumstances, the
composition of the solution will remain constant and so will the temperature of
the solution although it will exceed 100 °C.

45.

46.During the course of the distillation, the water vapor which distilled was
initially at the temperature of the solution. Suspending a thermometer above
this solution will record the temperature of the escaping vapor. As it departs
from the solution, the temperature of the vapor will cool by collisions with the
surface of vessel until it reaches 100 °C. Cooling below this temperature will
cause most of the vapor to condense to a liquid. If cooling to 20 °C occurs in
the condenser of a distillation apparatus, then by using the appropriate
geometry as shown in Figure 3, it would be possible to collect nearly all of the
liquid. The only vapor that would be lost to the environment would be that
small amount associated with the vapor pressure of water at 20 °C. Since the
vapor pressure of water at 20 °C is roughly 2.3 kPa, then 2.3/101.325 or 0.023
would be the fraction of water that would not condense and would pass out of
the condenser. This is why the distillate is frequently chilled in an ice bath
during the distillation.

47.

48.The distillation of a volatile material from non-volatile is one practical use of


distillation which works very well. However, often there may be other
components present that although they may differ in relative volatility, are
nevertheless volatile themselves. Let's now consider the two component system
you will be using in the distillations you will perform in the laboratory,
cyclohexane and methylcyclohexane. The vapor pressures of these two
materials in pure form are given in Table 1. As you can see from this table,
although cyclohexane is more volatile than methylcyclohexane, the difference
in volatility between the two at a given temperature is not very great. This
means that both materials will contribute substantially to the total vapor
pressure exhibited by the solution if the distillation is carried out at 1
atmosphere. The total pressure, PT, exerted by the solution against the
atmosphere according to Dalton's Law of partial pressures, equation 2, is
simply the sum of the observed vapor pressures of cyclohexane, , and
methylcyclohexane, :

49.

50.PT = + .2

51.

52.As before, boiling will occur when the total pressure, P T , equals an
atmosphere. However since we have two components contributing to the total
pressure, we need to determine the relative contributions of each. Again we can
use Raoult's Law but we need more information about the system before we
can do so. In particular we need to know the composition of the solution of
cyclohexane and methylcyclohexane. For ease of calculation, let's assume that
our original solution has equal molar amounts of the two components. What we
would like to determine is whether it would be possible to separate cyclohexane
from methylcyclohexane by distillation. By separation, we would like to
determine if it would be possible to end up with two receiver flasks at the end
of the experiment that would contain mainly cyclohexane in one and mainly
methylcyclohexane in the other. It is clear that at some point we will need to
intervene in this

53.Table 1. Vapor pressures of cyclohexane and methyl cyclohexane as a function


of temperature.

54.

cyclohexane methylcyclohexane

T/K P/kPa T/K P/kPa

300 14.1 300 6.7

305 17.6 305 8.5


310 21.7 310 10.6

315 26.5 315 13.2

320 32.2 320 16.2

325 38.8 325 19.8

330 46.5 330 24.0

335 55.3 335 28.9

340 65.4 340 34.6

345 77.0 345 41.2

350 90.0 350 48.7

354 101.3 354 55.4

360 121.3 360 66.9

362 128.5 362 71.1

365 139.9 365 77.9

370 160.5 370 90.2

373 174.0 373 101.3

380 208.8 380 119.3


385 236.7 385 136.4

390 267.3 390 155.3

395 300.9 395 176.2

400 337.5 400 199.1

55.

56.experiment. Otherwise, if we were to collect the entire contents of the original


distilling flask, called the pot, into one receiver flask, we would end up with the
same composition as we started. Initially the mole fractions of both
cyclohexane and methylcyclohexane are 0.5. From Raoult's Law (equation 1),
Dalton's Law (equation 2) and the information in Table 1, we can estimate that
boiling will occur at approximately 362 K when the total pressure of the two
components equals one atmosphere or 101.3 kPa.:

57.PT = + = 0.5(128.5 kPa) + 0.5(71.1 kPa)  101.3 kPa

58.

59.The first thing that we should note is that the initial boiling point is higher than
the lowest boiling component and lower than the highest boiling component.
Next, we should inquire about the composition of the vapor. Is the composition
of the vapor the same as the initial composition of the pot or is it enriched in
the more volatile component? If the composition of the vapor is the same as
that of the original mixture, then distillation will not be successful in separating
the two components. However, we should ask, "What is the composition of the
vapor?" Before the vapor is cooled and condenses on the condenser, we can
treat the vapor as an ideal gas. Recalling that: PV = nRT, where P is the
pressure of the gas or vapor, V is the volume it occupies, n is the number of
moles of gas, R is the Gas Constant (0.0821 L. atm.K -1.mol-1) and T is the
temperature, we can determine the composition of the vapor by taking
advantage of the following factors. First we note that:

60.

61.PTV = nTRT so that V = ncRT and V = nmRT where nT refers to the


total number of moles, since ( + ) = (nc + nm)RT.

62.If the total vapor can be treated as an ideal gas, then according to Dalton's Law,
so can each of the components. Since the two components are in thermal
contact and are distilling together, we can expect them to be at the same
temperature. We don't necessarily know the volume of the container, but since
it is assumed that the volumes of the molecules are very small in comparison to
the total volume the gas occupies, whatever the value of V, it is the same for
both components. This means we can establish the following ratio:

63.

64. / = nc/nm

65.

66.which in turn allows us to determine the composition of the vapor from the
observed partial pressures of the two components. If we use the experimental
values found in Table 1, we conclude that the composition of the vapor is 1.8/1,
and is indeed enriched in the more volatile component.

67.

68.This simple treatment allows us to understand the principles behind distillation.


However it is important to point out that distillation is far more complex than
our simple calculation indicates. For example, we just calculated the
composition of the vapor as soon as the solution begins to boil and we have
correctly determined that the vapor will be enriched in the more volatile
component. This means that as the distillation proceeds, the pot will be
enriched in the less volatile component. Since the composition of the pot will
change from the initial 1:1 mole ratio and become enriched in the less volatile
component; the new composition in the pot will introduce changes in the
composition of the vapor. The composition of the vapor will also change from
the initial ratio we just calculated to a new ratio to reflect the new composition
of the pot. The consequences of these changes are that the temperature of both
the pot and the distillate will slowly increase from the initial value to a value
approaching the boiling point and composition of the less volatile component.
If we are interested in separating our mixture into components, we are left with
the task of deciding how much material to collect in each receiver and how
many receivers to use. Obviously this will depend on the quality of separation
we are interested in achieving. Generally, the more receivers we use, the less
material we will have in each. It is possible to combine fractions that differ very
little in composition but this requires us to analyze each mixture. While it is
possible to do this, in general, we really want to end with three receivers, one
each enriched in the two components of our mixture and a third that contains a
composition close to the initial composition.

69.

70.It is difficult to describe how much material to collect in each receiver since the
volume collected will depend on the differences in the boiling points of the
components. As a general rule, the receiver should be changed for every 10 °C
change in boiling point. Each fraction collected can be analyzed and those with
compositions similar to the initial composition can be combined. The main
fractions collected can then be fractionated a second time if necessary.
71.The experiment we have just discussed is called a simple distillation. It is an
experiment that involves a single equilibration between the liquid and vapor.
This distillation is referred to as involving one theoretical plate. As you will
see, it is possible to design more efficient distillation columns that provide
separations on the basis of many theoretical plates. Before discussing these
columns and the advantages offered by such fractionating columns, it is
important to understand the basis of the advantages offered by columns with
many theoretical plates. The following is a simplified discussion of the process
just described involving a column with more than one theoretical plate.

72.

73.Fractional Distillation

74.

75.We have just seen that starting with a composition of 1:1, cyclohexane:
methylcyclohexane, the composition of the vapor was enriched in the more
volatile component. Suppose we were to collect and condense the vapor and
then allow the resulting liquid to reach equilibrium with its vapor. Let’s call this
liquid, liquid 2. The properties of liquid 2 will differ from the original
composition in two ways. First, since the composition of liquid 2 is higher in
cyclohexane than the initial one; the temperature at which liquid 2 will boil will
be lower than before (what is the approximate boiling point of a 1.8/1 mixture
of cyclohexane/methylcyclohexane? see Table 1). In addition, the composition
of the vapor, vapor 2, in equilibrium with liquid 2 will again be enriched in the
more volatile component. This is exactly what happened in the first
equilibration (first theoretical plate) and this process will be repeated with each
new equilibration. If this process is repeated many times, the vapor will
approach the composition of the most volatile component, in this case pure
cyclohexane, and the liquid in the pot will begin to approach the composition of
the less volatile component, methylcyclohexane. In order for this distillation to
be successful, it is important to allow the condensed liquid which is enriched in
the less volatile component relative to its vapor, to return to the pot. In a
fractional distillation, the best separation is achieved when the system is kept as
close to equilibrium as possible. This means that the cyclohexane should be
removed from the distillation apparatus very slowly. Most fractional distillation
apparati are designed in such a way as to permit control of the amount of
distillate that is removed from the system. Initially the apparatus is set up for
total reflux, (i.e. all the distillate is returned back to the system). Once the
distillation system reaches equilibrium, a reflux to takeoff ratio of about 100:1
is often used (about 1 out of every 100 drops reaching the condenser is
collected in the receiver).

76.

77.A column which allows for multiple equilibrations is called a fractionating


column and the process is called fractional distillation. An example of a
fractionating column is shown in Figure 4. Each theoretical plate is easy to
visualize in this column. The column contains a total of 4 theoretical plates and
including the first equilibration between the pot and chamber 1 accounts for a
total of 5 from pot to receiver. As you might expect, a problem with this
column is the amount of liquid that is retained by the column. Many other
column designs have been developed that offer the advantages of multiple
theoretical plates with low solvent retention. Typical spinning band columns
often used in research laboratories offer fractionating capabilities in the
thousand of theoretical plates with solute retention of less than one mL.
Commercial distillation columns have

78.

79.

80.Figure 4. A fractionating column which contains four chambers, each with a


center opening into the chamber directly above. The vapor entering the first
chamber cools slightly and condenses, filling the lower chamber with liquid. At
equilibrium, all chambers are filled with distillate. A portion of the liquid
condensing in the first chamber is allowed to return to the pot. The remaining
liquid will volatilize and travel up the column condensing in the second
chamber and so on. As discussed in the text, the composition of the vapor at
each equilibration is enriched in the more volatile component. The heat
necessary to volatilize the liquid in each chamber is obtained from the heat
released from the condensing vapors replacing the liquid that has been
removed. The vacuum jacket that surrounds the column ensures a minimum of
heat loss.

81.

82.been designed for gasoline refineries that are multiple stories high and are
capable of separating compounds with boiling points that differ by only a few
degrees.

83.In addition to performing a fractional distillation at one atmosphere pressure, it


is also possible to conduct fractional distillations at other pressures. This is
often avoided when possible because of the increased difficulty and expense in
maintaining the vacuum system leak free.

84.

85.Steam Distillation

86.

87.The concentration and isolation of an essential oil from a natural product has
had a dramatic impact on the development of medicine and food chemistry. The
ability to characterize the structure of the active ingredient from a natural
product has permitted synthesis of this material from other chemicals, resulting
in a reliable and often cheaper sources of the essential oil. The process often
used in this isolation is called steam distillation. Steam distillation is an
important technique that has significant commercial applications. Many
compounds, both solids and liquids, are separated from otherwise complex
mixtures by taking advantage of their volatility in steam. A compound must
satisfy three conditions to be successfully separated by steam distillation. It
must be stable and relatively insoluble in boiling water, and it must have a
vapor pressure in boiling water that is of the order of 1 kPa (0.01) atmosphere.
If two or more compounds satisfy these three conditions, they will generally
not separate from each other but will be separated from everything else. The
following example, expressed as a problem, illustrates the application of steam
distillation:

88.

89.Suppose we have 1 g of an organic compound present in 100 g of plant material


composed mainly of macromolecular material such cellulose and related
substances. Let's assume that the volatile organic material has a molecular
weight of 150 Daltons, a vapor pressure of 1 kPa and is not soluble in water to
an appreciable extent. Examples of such materials characterized by these
properties include eugenol from cloves, cinnamaldehyde from cinnamon bark
or cuminaldehyde from cumin seeds. How much water must we collect to be
assured we have isolated all of the natural oil from the bulk of the remaining
material?

90.We can simplify this problem by pointing out that the organic material is not
appreciably soluble in water. We know from previous discussions that boiling
will occur when the total pressure of our system equals atmospheric pressure.
We can also simplify the problem by assuming that the essential oil in not
appreciably soluble in the macromolecular material. While in reality this does
not have to be correct, this assumption simplifies our calculation. Boiling of our

91.

92.PT = + ; PT =  water +  org

93.

94.mixture will occur close to 100°C. Remember that very little oil is soluble in
water which makes the mole fraction of water near unity. Similarly for the our
volatile oil, its mole fraction is also close to one according to our assumption.
The total pressure, PT, is the sum of the vapor pressure of water, 100 kPa, and
the essential oil, , 1 kPa. Boiling will occur very close to the boiling point
of pure water. Treating the water vapor and the organic vapor which are
miscible as ideal, the PV ratio for both vapors is given by the following:

95.

96.PwaterV/ PorgV = nwaterRT/norgRT;

97.

98.Pwater/ Porg = nwater/norg and

99.nwater = wtwater/18; norg = 1/150; rearranging:

100.

101. wtwater = (100/1)(18/150) = 120 g water or 120 mL

102.

103. Our calculation suggests that we can be assured that most of the 1 g of
the organic mater has been transferred by the steam if we condense and collect
120 mL of water. The basis of the separation by steam distillation is that while
the water and organic condensed phases are immiscible, the vapors of both are
miscible. Once condensed, the two separate again allowing for an easy
separation. As noted above, both liquids and solids can be distilled by steam.

104.

105.
Theory of Distillation Part 2 (Laboratory Manual)

By : James W Zubrick
Email: j.zubrick@hvcc.edu

What Does It All Mean?


Getting back to the temperature-mole fraction diagram (Figure 140), suppose you start
with a mixture such that the mole fractions are as follows: isobutyl alcohol, 0.60, and
isopropyl alcohol, 0.40. On the diagram, that composition is point A at a room
temperature of 20 °C. Now you heat the mixture and you travel upwards from point A to
point B; the liquid has the same composition, it’s just hotter.

At 95 ° C, point B, the mixture boils. Vapor, with the composition at point C comes
flying out of the liquid (the horizontal line tying the composition of the vapor to the
composition of the liquid is the liquid—vapor tie line.), and this vapor condenses (Point
C to point D), say, part of the way up your distilling column.
Look at the composition of this new liquid (Point E). It is richer in the lower-boiling
component. The step cycle B-C-D represents one distillation.

If you heat this new liquid that’s richer in isopropyl alcohol (Point D), you get vapor
(composition at Point G along a horizontal tie-line) that condenses to liquid H. So step
cycle D-G-H is another distillation. The two steps represent two distillations.
Temperature vs. Mole Fraction Calculations

Nor
mal
Compound
BP
(C)
2-Propanol 82.3 1440 torr
2-Methyl-l- 108.
570 torr
propanol 2
Data from Moore
(3rd. ed.)
2-
Pro
9515.73 cal/mol
pan
ol
2
-Me
thyl
- 1- 10039.70 cal/mol
pro
pan
ol
Comparison of Data from Moore and Calculated Data (T= 100 (C))

Moore Calc.
Vapor pressures: 2-Propanol 1440 1440.05
(torr) 2-Methyl-l-propanol 570 570.02
Mole fraction: 2-Propanol 0.219 0.2184
(in liquid) 2-Methyl-2-propanol 0.781 0.7861
Mole fraction: 2-Propanol 0.415 0.4137
(in vapor) 2-Methyl-2-propanol 0.585 0.5863
Mole Fraction of 2-Propanol Liquid Data

%
XC X Dif
Dif
alc Lit. f
f

-
0.9 0.9 -.0
1 83.2 0.2
458 485 027
86
-
0.8 0.8 -.0
2 85.4 0.7
213 275 062
48
-
0.7 0.7 -.0
3 86.9 0.3
425 450 025
31
0.6 0.6 0.0 2.5
4 88.7
540 380 160 05
0.5 0.5 0.0 1.5
5 90.9
539 455 084 40
0.3 0.3 0.0 3.9
6 95.8
591 455 136 49
0.2 0.2 0.0 1.3
7 99.9
215 185 030 57
0.1 0.1 0.0 3.0
8 103.4
191 155 036 96
-
0.0 0.0 -.0
9 106.2 1.5
458 465 007
07
Mole Fraction of 2-Propanol Vapor Data
%
XC X Dif
Dif
alc Lit. f
f

-
0.9 0.9 -.0
1 83.2 0.2
785 805 020
01
-
0.9 0.9 -.0
2 85.4 0.7
228 295 067
23
-
0.8 0.8 -.0
3 86.9 0.6
820 875 055
18
0.8 0.8 0.0 0.6
4 88.7
300 245 055 63
0.7 0.7 0.0 0.4
5 90.9
615 580 035 60
0.5 0.5 0.0 0.6
6 95.8
880 845 035 00
-
0.4 0.4 -.0
7 99.9 2.0
182 270 088
63
0.2 0.2 0.0 0.9
8 103.4
533 510 023 36
-
0.1 0.1 -.0
9 106.2 4.4
070 120 050
33
Fig. 140 Temperature-mole fraction diagram for the isobutyl-isoropyl alcohol
systems.
Much of this work was carried out using a special distilling column called a bubble-
plate column (Fig. 141). Each plate really does act like a distilling flask with a very
efficient column, and one distillation is really carried out on one physical plate. To
calculate the number of plates (separation steps, or distillations) for a bubble-plate
column, you just count them!
Unfortunately, the fractionating column you usually get is not a bubble-plate type. You
have an open tube that you fill with column packing (see “Class 3: Fractional
Distillation”) and no plates. The distillations up this type of column are not discreet, and
the question of where one plate begins and another ends is meaningless. Yet, if you use
this type of column, you do get a better separation than if you used no column at all. It’s
as if you had a column with some bubble-plates. And if your distilling column separates
a mixture as well as a bubble-plate column with two real plates, you must have a
column with two theoretical plates.
You can calculate the number of theoretical plates in your column if you distill a
two-component liquid mixture of known composition (isobutyl and isopropyl alcohols
perhaps?), and collect a few drops of the liquid condensed from the vapor at the top of
the column. You need to determine the composition of that condensed vapor (usually
from a calibration curve of known compositions versus their refractive indices [see
Chapter 22, "Refracto-metry"]), and you must have the temperature-mole fraction
diagram (Fig. 140).
Fig. 141 A bubble-plate fractionating column.
Suppose you fractionated that liquid of composition A, collected a few drops of the
condensed vapor at the top of the column, analyzed it by taking its refractive index, and
found that this liquid had a composition corresponding to point J on our diagram. You
would follow the same path as before (B-C-D, one distillation; D-G-H, another
distillation) and find that composition J falls a bit short of the full cycle for distillation #2.
Well, all you can do is estimate that it falls at, say, a little more than half of the way
along this second tie-line, eh (Point K)? OK then. This column has been officially
declared to have 1.6 theoretical plates. Can you have tenths of plates? Not with a
bubble-plate column, but certainly with any column that does not have discrete
separation stages.
Now you have a column with one-point-six theoretical plates. “Is that good?” you
ask. “Relative to what,” I say. If that column is six feet high, that’s terrible. The Height
Equivalent to a Theoretical Plate (HETP) is 3.7feet/ plate. Suppose another column also
had 1.6 theoretical plates, but was only 6 inches (0.5 ft) high. The HETP for this column
is 3.7in/plate, and if it were 6 feet high, it would have 19 plates. The smaller the HETP,
the more efficient the column is. There are more plates for the same length.
One last thing. On the temperature-mole fraction diagram, there’s a point F I haven’t
bothered about. F is the grade you’ll get when you extend the A-B line up to cut into the
upper curve and you then try to do anything with this point. I’ve found an amazing
tendency for some folk to extend that line to point F. Why? Up the temperature from A to
B and the sample boils. When the sample boils, the temperature stops going up. Heat
going into the distillation is being used to vaporize the liquid (heat of vaporization, eh?)
and all you get is a vapor, enriched in the lower-boiling component, with the composition
found at the end of a horizontal tie-line.

Reality Intrudes I: Changing Composition


To get the number of theoretical plates, we fractionally distilled a known mixture and
took off a small amount for analysis, so as not to disturb things very much. You,
however, have to fractionally distill a mixture and hand in a good amount of product, and
do it within the time limits of the laboratory.
So when you fractionally distill a liquid, you continuously remove the lower boiling
fraction from the top of the column. And where did that liquid come from? The boiling
liquid at the bottom of the column. Now if the distillate is richer in the lower boiling
component, what happened to the composition of the boiling liquid? I’d better hear you
say that the boiling liquid gets richer in the higher-boiling component (Fig. 142).

Fig. 142 Changing composition as the distillation goes on.


So as you fractionally distill, not only does your boiling liquid get richer in the higher-
boiling component, so also does your distillate, your condensed vapor. Don’t worry too
much about this effect. It happens as long as you have to collect a product for
evaluation. Let your thermometer be your guide, and keep the temperature spread less
than 2°C.

Reality Intrudes II: Nonequilibrium Conditions


Not only were we forced to remove a small amount of liquid to accurately
determine the efficiency of our column, we had to do it very slowly. This allowed the
distillation to remain at equilibrium. The throughput, the rate at which we took material
out of the column, was very low. Of all the molecules of vapor that condensed at the top
of the column, most fell back down the column; few were removed. A very high reflux
ratio. With an infinite reflux ratio (no throughput), the condensed vapor at the top of the
column is as rich in the lower-boiling component as it’s ever likely to get in your setup.
As you remove this condensed vapor, the equilibrium is upset, as more molecules rush
in to take the place of the missing. The faster you distill, the less time there is for
equilibrium to be reestablished—less time for the more volatile components to sort
themselves out and move to the top of the column. So you begin to remove higher-
boiling fractions as well, and you cannot get as clean a separation. In the limit, you
could remove condensed vapor so quickly that you shouldn’t have even bothered using
a column.

Reality Intrudes III: Azeotropes


Occasionally, you’ll run across liquid mixtures that cannot be separated by fractional
distillation. That’s because the composition of the vapor coming off the liquid is the
same as the liquid itself. You have an azeotrope, a liquid mixture with a constant boiling
point.
Go back to the temperature-mole fraction diagram for the isopropyl alcohol-isobutyl
alcohol system (Fig. 140). The composition of the vapor is always different from that of
the liquid, and we can separate the two compounds. If the composition of the vapor is
the same as that of the liquid, that separation is hopeless. Since we’ve used the notions
of an ideal gas in deriving our equations for the liquid and vapor compositions
(Clausius-Clapyron, Dalton, and Raoult), this azeotropic behavior is said to result from
deviation from ideality, specifically deviations from Raoult’s Law. Although you might
invoke certain interactive forces in explaining nonideal behavior, you cannot predict
azeotrope formation a priori. Very similar materials form azeotropes (ethanol-water).
Very different materials form azeotropes (toluene-water). And they can be either
minimum-boiling azeotropes or maximum-boiling azeotropes.

Minimum-Boiling Azeotropes
The ethanol-water azeotrope (95%ethanol-5%water) is an example of a minimum
boiling azeotrope. Its boiling point is lower than that of the components (Fig. 143). If
you’ve ever fermented anything and distilled the results in the hopes of obtaining 200
proof (100%) white lightning, you’d have to content yourself with getting the azeotropic
190 proof mixture, instead. Fermentation usually stops when the yeast die in their own
15% ethanol solution. At room temperature, this is point A on our phase diagram. When
you heat the solution, you move from point A to point B and, urges to go to point F
notwithstanding, you cycle through distillation cycles B-C-D and D-E-G, and, well, guess
what comes off the liquid? Yep, the azeotrope. As the azeo-trope comes over, the
composition of the boiling liquid moves to the right (it gets richer in water), and finally
there isn’t enough ethanol to support the azeotropic composition. At that point, you’re
just distilling water. The process is mirrored if you start with a liquid that is > 95%
ethanol and water. The azeotrope comes off first.
Fig. 143 Minimum-boiling ethanol-water azeotrope.

N E X T P O S T: Theory of Distillation Part 3 (Laboratory Manual)

P R E V I O U S P O S T: Theory of Distillation Part 1 (Laboratory Manual)


 R ELATED L INKS

o Organic Chemistry Laboratory Survival Manual


 Safety First, Last and Always (Laboratory Manual)
 Keeping a Notebook (Laboratory Manual)
 Interpreting a Handbook Part 1 (Laboratory Manual)
 Interpreting a Handbook Part 2 (Laboratory Manual)
 Jointware Part 1 (Laboratory Manual)
 :: SEARCH WWH ::

Вам также может понравиться