Sei sulla pagina 1di 13

Adamson University

College Of Engineering
Chemical Engineering Department

Unit Operations Lab 2


Experiment No. 2
DRYING CURVES
Submitted by:
Group 8
Briones, Neil Ryan V.
Paguia, Nikka C.
Rodriguez, John Paul
Magleo, Gelo
Sangalang, Joey
Modina, Ma. Betina V.

Submitted to:
Engr. Rubi Rudgi

January 19, 2016

Experiment No. 2
DRYING CURVES
ABSTRACT

The objective of this work is to analyze the possibility of applying the drying curves
generalization methodology to the conductive/convective hot plate drying of cellulose. The
experiments were carried out at different heated plate temperatures and air velocities over the
surface of the samples. This kind of approach is very interesting because it permits comparison
of the results of different experiments by reducing them to only one set, which can be divided
into two groups: the generalized drying curves and the generalized drying rate curves.

TRANSMITTAL LETTER

October 6, 2015
Engr. Albert Evangelista
Chemical Engineering Department
Adamson University
Ermita, Manila

Engr. Evangelista:
In compliance with the fulfillment of the requirements on the subject Unit Operations Lab
2, the group would like to present this experiment report entitled DRYING CURVES in
accordance with your instructions. The main purpose of this experiment report is to
determine heat flow rate through the bare and lagged pipes, to determine the thermal
conductivity of lagging material by assuming the heat input to be that heat flow rate through
lagged pipe and to determine the efficiency of insulating materials. We hope that this
experiment report will meet you approval.

Respectfully Yours,
Briones, Neil Ryan V.
Paguia, Nikka C.
Rodriguez, John Paul
Magleo, Gelo
Sangalang, Joey
Modina, Ma. Betina V.
I.

Objectives

1. To produce drying and drying rate curves for a wet solid being dried with air of
fixed temperature and humidity
2. To determine the critical and equilibrium moisture contents, bound and unbound
moisture of the wet solid being dried.
II.

Materials/Equipment Needed
1. Tray Drier
2. Sieve shaker
3. Sand sieved to approximately 500 microns
4. 2-Sling Psychrometers
5. 7-Thermometers
6. Weighing Instrument
7. Oven (if necessary)
8. Stop watch
9. Water
10. Container/pale
11. Anemometer/Wind Speed meter
12. Boiler

III.

Equipment Set up

IV.

Theory
In the process of drying heat is necessary to evaporate moisture from the grain and a
flow of air is needed to carry away the evaporated moisture. There are two basic
mechanisms involved in the drying process; the migration of moisture from the interior
of an individual grain to the surface, and the evaporation of moisture from the surface
to the surrounding air. The rate of drying is determined by the moisture content and the
temperature of the grain and the temperature, the (relative) humidity and the velocity of
the air in contact with the grain.
Drying only takes place if the wet material contains more moisture than the equilibrium
value for its environment. The earliest ideas on convective drying implied that liquid
moisture diffuses to the exposed surface of a wet body where it evaporates, the vapor
diffusing through the boundary layer into the bulk of the surrounding air. This view is
clearly unsatisfactory, except for drying of homogeneous materials in which the
moisture is effectively dissolved. Mechanisms of moisture movement are generally
more complex. Most materials are composed of sub entities, such as particles and fibers,
which may be loose or held in some kind of matrix. The number and nature of the voids
between these entities and the pores within them govern the quantity of moisture
retained and the extent of bonding to the solids. If the openings form a capillary

network, the material is said to be capillary-porous. A capillary-porous material may


be nonhygroscopic: that is, the moisture held within the body exerts its full vapor
pressure. This is a limiting case found in some coarse, nonporous mineral aggregates.
Moisture is simply trapped between the particles. As the space between the particles
becomes more confined, vapor pressure is lowered according to the Kelvin equation

where pk is the capillary-vapor pressure, p0 is the saturation vapor pressure, is the


surface tension, is the molar liquid volume and dp is the capillary size.
Hygroscopicity may be related to this capillary condensation in the voids, but normally
it stems from the structure of the primary entities with their finer passageways and their
ability to hold moisture in a variety of ways. Material may not be simply capillaryporous, but be composed of complex arrangements of capillaries, vessels and cells, as
seen with materials of vegetable origin. Some of these may be described as being both
capillary-porous and colloidal, composed of a matrix colloidal by nature but developing
a pore structure upon drying. Colloidal slurries of inorganic particles lose volume as
moisture is driven off until the mass becomes consolidated. Colloidal material of
biological origin, by contrast, does not shrink until condensed moisture in the
intermiscellar spaces has been emptied. This condition is called the fibre-saturation
point in woody material drying. Thereafter, as cellular moisture is expelled, the stuff
shrinks as the cells shrivel.
The ratio of the moisture-vapor pressure to the saturation value at the same temperature
is called relative humidity . The lower the relative humidity, the more strongly is the
moisture bound to the host material. The free energy G required to release unit molal
quantity of this moisture is given by

for an isothermal, reversible process without change of composition. Isothermal


variation of the equilibrium moisture content (which is a function of this free-energy
change) with relative humidity yields a moisture isotherm, and it is the desorption
isotherm which is of interest in drying. Moisture isotherms are normally sigmoid in
shape when plotted over the whole relative humidity range, but often a simple

exponential expression can be fitted over a more limited range at higher relative
humidities. The general shape of the isotherm reflects the nature of the moist material,
as illustrated in Figure 3 An exception to this kind of behavior is that of inorganic
crystalline solids which have multiple hydrates. With these materials, relative humidity
falls in stepwise fashion with loss of moisture as each hydrate disappears.

Figure 3. Variation of relative humidity with equilibrium moisture content for


different materials. After Keey (1978).
When pores in a solid are of molecular size, moisture can only be held therein
by volume filling such that the adsorbate, although highly compressed, is not considered
to be a separate phase. The chemical potential of the adsorbent varies with the amount
adsorbed, unlike the behavior at higher moisture contents when a separate adsorbate
phase existswith equilibrium between phasesand the chemical potential remains
invariant. The fractional filling of these micropores is a complex exponential function of
free-energy sorption, RT ln. With larger pore sizes, moisture can sorb in molecular
layers on the host material. Consideration of multimolecular adsorption leads to the
equation

for the equilibrium moisture content X at a relative humidity , where X 1 is the


moisture content for a complete monolayer, k is the exp ( H/RT), where H is the
constant enthalpy difference in adsorptive heats between the first and successive
molecular layers of moisture, and C is a coefficient. When k is zero, the equation
reduces to that for monomolecular adsorption. Over the range 0 < k < 1, this equation
can describe the moisture-sorption behavior of materials, which appears to reach a finite
moisture content as relative humidity approaches unity. This quantity is sometimes
known as the maximum hygroscopic moisture content. This three-coefficient equation

(with C, k and X1 as the adjustable parameters) has been tested for sorption of water
vapor on 29 materials at room temperature over a wide range of relative humidity (from
0.07 to 0.97) and for some of these materials, over a narrower range of temperatures
between 45 and 75%C [Jaafar and Michalowski (1990)]. In most cases, experimental
data could be fitted to within 8% up to a relative humidity of 0.7, and in some
instances over the whole humidity range. To cope with the hygroscopic behavior at high
relative humidities with colloidal material, which swells with increasing moisture
content, Schuchmann et al. (1990) have recommended that ln (1 ) be chosen as the
dependent variable rather than y itself in the correlation. It is unwise to extrapolate
sorption correlations beyond their tested range of relative humidities, owing to changes
in hygroscopic behavior at extremes in this range compared with that at intermediate
values.
The manner in which a material dries out depends not only on its structure but also on
its physical form. The drying of small wood chips is controlled essentially by moisturevapor transport through the boundary layer; veneers and thin slats of the same wood by
the dry fraction of the exposed surface; while the drying of board timber, by the internal
moisture-transport mechanisms within the timber itself. Early experiments on drying
materials in sample trays in an air stream have noted that initially, the drying rate was
almost the same as that of a free liquid surface under the same conditions and remained
relatively constant as the material dried out [Keey (1972)]. This period of drying is
followed by one in which the drying rates fell off sharply as moisture content was
reduced to the equilibrium value even though the drying conditions remain unchanged.
This marked difference in behavior has led to the division of drying into the constantrate period and falling-rate period, respectively. The "knee" in the drying curve
between these two periods is known as the critical point. Sometimes, these periods are
referred to as unhindered dryingand hindered drying, respectively, to indicate whether
the material itself plays a controlling role in restricting moisture loss. Appearance of the
initial period can be masked by the induction effects at the start of drying as a moist
solid warms up or cools to a dynamic equilibrium temperature, which is the Wet-bulb
Temperature, if the surface is only heated convectively. This surface temperature is
maintained as long as the surface is sufficiently wet to effectively saturate it. An
example of a drying curve is shown in Figure 4.

Figure 4. An example of a drying curve.


The reasons for the appearance of a drying period of constant, or near-constant, drying
rates are complex, particularly as it is unlikely that there exists a film of liquid moisture
at the surface except in rare circumstances [van Brakel (1980)]. Indeed, a constant-rate
period can be observed if the dimensions of the wet and dry patches on the surface are
sufficiently small compared with the thickness of the boundary layer. The requirement is
that moisture-vapor pressure at the surface maintain the saturation value at the mean
surface temperature, and thus the rate of moisture loss over unit exposed surface (

is:

where pG is the partial pressure of the moisture vapor in the bulk of the gas and p S is the
value of the gas adjacent to the moist surface. In drying calculations, it is more useful to
use humidities (ratios of the mass of moisture vapor to that of dry gas), and the above
expression transforms to

where y is a mass transfer coefficient based on the humidity difference; is the


humidity potential coefficient which effectively "corrects" for the introduction of a
linear humidity driving force [Keey (1978)]; and YS and YG are the humidities at the
surface and in the bulk of the gas, respectively. The falling-rate period begins when
moisture movement within the solid can no longer maintain the rate of evaporation, or if
moisture content at the surface falls below the maximum hygroscopic value and the
partial pressure of the moisture vapor also dwindles. Clearly, a constant-rate period may

never be observed in colloidal material drying that reaches unit relative humidity only at
very high moisture contents.
To a first approximation, the drying kinetics in the falling-rate period may be regarded
as of the first-order, and thus the drying rate is directly proportional to the difference
between the mean moisture content of the wet material (X) and its equilibrium value
(Xe):

Integration of this equation yields the time t to dry from a moisture content of X 1 to
X2:

where a is a coefficient which may be proportional to a moisture "diffusivity." However,


the validity of this equation is no test that moisture movement is by diffusion as this
expression correlates drying time for a number of materials dried in static and fluidized
beds, in which diffusional processes are unlikely [van Brakel (1980)]. The appearance
of first-order kinetics is sometimes referred to as the regular regime of drying.
There have been a number of attempts to provide a fundamental theoretical framework
to describe drying and the development of moisture-content and temperature profiles in
materials being dried. These include the use of irreversible thermodynamics to define
the appropriate transport potentials [Luikov (1966)]; theories based on moisture-vapor
diffusion and capillary transport of liquid in porous media [Krischer and Kast (1978)];
and volume-averaging the equations of continuity, mass and energy conservation which
apply to each of the discontinuous phases within a moist porous body [Whitaker
(1980)]. While these approaches have found some limited success in describing heat and
mass transfer processes under certain idealized conditions, these theories are limited in
application by the assumptions made to get numerical solutions.
In practice, more empirical approaches have been found which are useful in describing
drying behavior of actual material under industrial conditions. One such method is
based on the concept of the characteristic drying curve [Keey (1978)]. This is a
generalized drying curve obtained from laboratory tests under constant drying
conditions with a sample material. It is a plot of the drying rate, normalized with respect

to its maximum value (mW) in the constant-rate period, against a characteristic moisture
contentdefined as the ratio of the freely evaporable moisture content (X X e) to the
free moisture content at the critical point (Xcr Xe). These nondimensional parameters
thus become

and

respectively. An example of a characteristic drying curve is shown in Figure 5.

Figure 5. An example of a drying curve.


A unique characteristic drying curve is only found if the ratio of the exposed surface to
the material volume is maintained constant or has a unique value at a given
characteristic moisture content if the material shrinks. Strictly speaking, such a curve is
only found at extremes of drying intensity, when the material's moisture content is
essentially uniform (low-intensity drying) or when there is a sharp front between driedout material close to the surface and a still-wet interior (high-intensity drying). In
practice, however, over the range of practical drying conditions, characteristic drying
curves appear when drying a variety of particulate and loose materials, provided the
particle size is below 20 mm [Keey (1992)]. There is normally no simple relationship
between the parameters f and , although there are some special cases. First-order

kinetics correspond to the identity f = , with drying of permeable materials


approximating this behavior. When drying is controlled by the dry fraction of exposed
surface (with thin sheets, for example), then f = 2/3. Drying of impermeable materials
like heartwood timber corresponds approximately to f = 2. If no critical point is
observed in the drying experiment, a characteristic curve can be drawn up based on
normalizing the moisture content with respect to the ostensible start of the falling-rate
period, provided a simple algebraic relationship can be fitted to the characteristic curve.
The particular advantage of the characteristic drying curve concept is that a simplified
equation can be written to describe the rate of drying at any place within a dryer if the
humidity potential (YW YG) is known, where YW is the saturation humidity at the wetbulb temperature:

Examples of the use of this expression to describe industrial drying processes are given
by Keey (1978, 1992).
The intensity of drying (I) is given by the ratio of the maximum unhindered drying rate
(mW) to the maximum moisture-transfer rate through the material assuming a diffusionlike process:

where D is a moisture diffusion coefficient, X0 is the initial moisture content and b is the
effective thickness or radius of the material. This suggests that the product mWb is a
useful property; it is called the flux parameter, F. If ln F is plotted against moisture
content X, a separate curve is found for each initial moisture content X 0 in the
penetration period as the moisture content and temperature profiles develop within the
material. In the regular regime, there is a common curve independent of initial moisture
content and drying flux.
Under certain circumstances, drying curves can appear with both negative and positive
gradients in drying rate as moisture is lost. The drying of layers of soluble dyestuffs can
show discontinuities due to the build up and cracking of surface crusts, particularly if
the crust is removed periodically. The drying of porous bodies containing a mixed

volatile solvent may also show periods of falling and rising rates due to selective
evaporation of the moisture with changes in relative volatility as the composition alters.

References to nikka:
http://www.thermopedia.com/content/711/
http://www.scielo.br/scielo.php?script=sci_arttext&pid=S010466322003000100015

paki-ayos nden ung part ng names =) HAHAHA

Potrebbero piacerti anche