Sunteți pe pagina 1din 13

Solid-Gas Separation

The separation of solids from gas streams is a common unit operation in production plants. Gas-
solid separation is important both as a device to recover product but also a key technology for
environmental control. An incredible array of process technologies (for example, cyclones, bag
houses, spray towers, venturi scrubbers) is available to accomplish the task of separating the
solids from gas.
This webinar first examines the concept of grade efficiency, a key concept for describing the
efficiency of a solids separation device. Second, it addresses the topic of selecting a particular
process technology for a given gas-solid separation problem. We also examine advantages and
trade-offs of the most common solid-gas separation devices. Lastly, the fundamental science
associated with the devices such as cyclones, bag houses (dust collectors), and some wet
scrubbing technologies are addressed.

Particulate solids are separated from gas streams for a variety of reasons. In some instances, the
interest is in the recovery of solids as a product, e.g., following a milling operation for which a
combined classification and separation is often required [Prasher ( 1987)]. In other cases, the
emission of fine particulate solids and dust from a unit operation may be excessive and, therefore,
a reduction and control of particulate level is required for the protection of subsequent process
equipment or for the environmental emission [Strauss (1975, and Theodore and Buonicore
(1976)]. This is illustrated inFigure 1 for the case of advanced power generation from coal,
where it is desirable to expand the hot gases in a gas turbine before exhaust to atmosphere to
enhance the efficiency of power generation. Typical particulate effluent from a pilot-scale
fluidized bed combustor is shown, together with upper limits of allowable emission to
atmosphere, and of the inlet dust tolerance of high performance gas turbines [Henry et al. (1982)].
In this section, an overview of the principles of gas-solids separation is given. Details of various
technologies are addressed in the next three sections.
The most commonly used techniques for the separation of particulate solids from gases
are inertial separators, Electrostatic Separators and Filters. The physical phenomena involved
in the separation of solids from gases are influenced by a number of important factors such as the
properties of the gas, and gas-particle and particle-particle interactions. These factors are briefly
reviewed here first before addressing the principles of various types of gas-solids separators.

Figure 1. Industrial process and environmental requirements for gas-solids separation.
Gas Properties
Over the range of pressures commonly used in industrial gas-solids separation processes,
departures from ideal gas behavior are negligible. Therefore the density of a gas at pressure P
and absolute temperature T can be approximated by
(1)

where Mw is the mean molecular weight of the gas, the molar volume, and the universal
gas constant.
Elementary kinetic theory gives a first approximation for the effect of temperature and pressure
on gas viscosity. Viscosityis predicted to be independent of pressure, and proportional to the
square root of absolute temperature.
In addition to these macroscopic properties, the Mean Free Path of gas molecules, , is of
interest for removal of fine particles. Elementary kinetic theory predicts that is inversely
proportional to density, so that
(2)

Gas-Particle Interaction
Almost any particle removal process requires the particles to migrate relative to the gas, so that
the drag of the gas on the particle is of prime interest. The particle Reynolds Number is defined
as
(3)

where d
p
is the particle diameter and u the velocity of the particle relative to the gas and the gas
viscosity. In most applications, Re
p
remains small. The drag force, F
D
, can then be estimated
from Stokes' Law with a correction to allow for "slip effects" which arise when the particle
diameter is comparable to the mean free path of gas molecules:
(4)

where C is the slip correction factor or "Cunningham coefficient" defined as
(5)

C can be estimated empirically from Davies' correlation (1945):
(6)

where Kn is the Knudsen Number:
(7)

Equation (6) is applicable to spherical particles, and Beard (1976), and Gift et al., (1978)
summarize modifications for nonspherical particles. Some representative values for C are given
by Clift et al., (1981). It is worth noting that, whereas C only departs significantly from unity for
submicron particles at ambient conditions and at elevated pressure, slip effects are significant for
particles several microns in diameter at elevated temperatures and ambient pressure.
For Reynolds numbers larger than, say, 0.1, Equation (4) is better replaced by the general form:
(8)

where C
D
is an empirical function of Re
p
Of the many forms suggested for C
D
(Re
p
), that due to
Schiller and Nauman (1933) is probably most accurate for Re
p
< 800:
(9)

When Re
P
is sufficiently large that Equation (8) must be used rather than Eq. (4), then slip
effects can be ignored, unless u is significant by comparison with the speed of sound in the gas.
However, this is not normally the case in industrial gas-solids separation processes.
For a particle settling freely under its own weight in a gas, the drag force counterbalances the
immersed weight of the particle. For low Re
p
, terminal velocity then follows from Equation
(4) as
(10)

where
p
is the particle density. Normally
p

g
, so that Eq. (1) can be written
(10a)

Increasing temperature, through its effect on gas viscosity, increases fluid-particle drag, reduces
settling velocity, and generally makes removal of particles from gases more difficult. The effect
of pressure is smaller, and acts through the slip effect so that it is only significant for particles
typically smaller than about 1 m; in general, the effect of increasing pressure is again to make
particle removal more difficult. Electrostatic precipitators can be an exception to this rule
because of the effect of pressure on the electrical properties of a gas. Increasing the pressure,
within the range of interest here, widens the gap between the corona-starting and sparkover
voltages. Therefore, the effect of increased drag on the particle at high pressure and high
temperature may be compensated or even overcome by increasing the electrostatic field intensity
(see below).
In certain types of filter the Brownian diffusivity of particles in the gas is of concern. This is
usually best evaluated by theStokes-Einstein equation [Clift et al. (1981)]:
(11)

where k
B
is Boltzmann's constant, 1.38062210
23
JK
1
. Taking the effect of temperature on
viscosity into account, for given particle size
(11a)

so that the Brownian diffusivity increases with increasing temperature. Pressure has a weaker
effect, through the slip correction factor, C, increasing pressure decreases D
AB
It follows from
the above that, if the efficiency of a gas-solids separation process is to be investigated, it is more
important to match the process temperature than pressure. Pressure is important for fine particles,
so that devices such as electrostatic precipitators intended to collect micron-sized particles would
be tested at process temperature and pressure; electrical properties also dictate that both pressure
and temperature should be matched.
Particle-Particle Interaction
In some devices for particle collection, most notably electrostatic precipitators and some types of
filter, it is essential for the collected particles to form relatively large agglomerates which can
then be removed easily. Particle-particle cohesive forces arising from electrostatic, van der
Waals and capillary effects are then important.
Electrostatic effects act over the longest range and may therefore be effective in separating
particles from gases. However, they can only contribute to the cohesiveness of a deposited
particulate if the constituent particles carry both positive and negative charges. Corny and Clift
(1984) have shown that fly-ash from fluidized combustion can carry such charges, and be highly
cohesive as a result. However, it is not known whether this phenomenon occurs more generally
nor, for example, whether ash which has been raised above the fusion temperature carries
significant charges. More generally, particles are likely to carry charges of the same sign, so that
interparticle Coulombic repulsion acts against cohesion. Leakage of this charge requires a
conductive path to an earthed surface, and this depends in particular on the surface
characteristics of the particles. If the particles have for some reasons low surface resistivity, then
electrical contact between particles can be good and the charge can leak away relatively rapidly.
This is particularly important in electrostatic precipitation.
The other two types of force act over shorter ranges, and thefore represent genuine cohesion.
Whereas electrostatic forces result from overall surplus of deficit of electrons, van der Waals
forces arise from the attraction between atomic or molecular dipoles. Because they result from
processes occurring on a molecular scale, van der Waals forces act over much shorter ranges
than electrostatic forces. The strength of van der Waals cohesion depends on the value of
the Hamaker constant, a characteristic property of the material comprising the particles. The
Hamaker constant is effectively independent of pressure, and the dependence on temperature is
not well understood but is probably weak. Hence the environment affects van der Waals
cohesion mainly through its effect on surface properties. Dahneke (1972) showed that local
deformation of the contacting surfaces can increase substantially the strength of interparticle
cohesion.
Capillary forces, which result from a liquid film on the surface of the particles, are orders of
magnitude stronger than van der Waals forces (see also Capillary Action). Figure 2 shows
schematically a "pendular bridge" between two idealized particles of diameter d
p
For zero
particle separation ( = 0) and a fully-wetting liquid (contact angle, = 0) the attractive force
between the particles is [Fisher (1926)]:
(12)

where is the angle defining the size of the bridge and is the tension at the gas/liquid interface.
According to Equation 12, the cohesive force increases as the liquid bridge becomes smaller, to
a limiting value for of d. This theoretical result is not applicable to real particles with
rough surface [Cheng (1970)]; for particles which contact at points of asperity the cohesive force
increases with liquid film thickness [Coughlin et al. (1982)].

Figure 2. Pendular liquid bridge between solid particles.
Principles of Gas-Solids Separation Devices
Inertial Separators cover devices in which the main property used in recovering particles is their
density, so that they are removed by centrifugal action. Deliberate changes in the direction of gas
flow causes the particles trajectories to deviate from the gas streamlines, thus concentrating and
separating the particles from the gas. Inertial separators are varied in design. Most separators in
this category use passive mechanical separation with induced centrifugal motion, such as
cyclones. There are however separators in which the centrifugal motion is induced by a rotating
propeller. These devices are often used for classification purposes, and are covered in Classifiers.
Inertial separators can be used in a wide range of pressure and temperature conditions. Their
performance is satisfactory for coarse particles, but as the particle size decreases much below 10
m the collection efficiency deteriorates very rapidly. Of the various types of design, the reverse
flow cyclone with tangential entry is the most common type of inertial separators due to its
compactness, simplicity of construction and operation, and high throughput. The inlet can be
made either truly tangential or wrapped around the body, commonly referred to as "tangential
entry" and "volute entry" or "scroll inlet", respectively. The former has a higher collection
efficiency, but at the expense of higher pressure drop. There are, of course, other designs using
axial entry with vanes installed within the cyclone to induce rotation, direct through-flow of gas
without reversing, and blowdown, where a small fraction of the gas is allowed to flow with the
collected solids [Strauss (1975)].
Performance analysis of cyclones has been approached on both phenomenlogical as well
mechanistic levels. On a phenomenological level, conventional dimensional analysis leads to the
definition of groups which can be used to "scale" cyclone performance for effects of size,
throughput and process conditions, see Abrahamson (1981). The collection efficiency, E, for
particles of diameter d
p
and density
p
from a gas of density
g
and viscosity is given by a
relationship of the form
(13)

while the pressure drop across the cyclone, P, follows a relationship with the general form
(14)

In Eqs. (13) and (14), is the volumetric flowrate of gas through the cyclone, and D is the
cyclone barrel diameter. Therefore represents a characteristic gas velocity, such as the
mean inlet gas velocity, as the inlet dimensions are fixed relative to the barrel diameter for a
given design. The two independent dimensionless groups in Eq. (13) can be regarded as
the cyclone Stokes number, St
c
, and Reynolds number, Re
c

(15)

(16)

The Stokes number describes the tendency of the particle trajectories to deviate from gas
streamlines, and the Reynolds number describes the gas flow condition within the cyclone. The
above analysis is only valid for low particle concentrations, where the probability of collection of
any individual particle is unaffected by other particles present. For industrial cyclones, the gas
flow in the cyclone is turbulent and Re
c
is very large. It is therefore commonly assumed that the
collection efficiency is not affected by the gas flow pattern, and the effect of Re
c
is neglected.
This leads to the commonly used scaling laws, see, for example, Strauss (1975):
(17)

(18)

Equations (17) and (18) have been shown to work well for low particle loadings and for
particles carrying low levels of electric charge [Giles (1982)]. However, for the cases where
particle/particle interactions are significant, including electrical effects, the above correlations
are not satisfactory and E is less sensitive to St
c

The effect of temperature and pressure is in principle reflected in the density and viscosity of the
gas. However, care must be taken in the use of Equation (17) in this case because temperature
and pressure can also modify Re
c
, which it can then alter the flow condition, thus producing an
effect which is not taken into account in Equation (17).
The alternative approach to analysis of cyclone performance is by mechanistic modelling, which
is necessary to account for the effect of pressure and temperature, as well as for predicting the
effects of changing cyclone geometry. Clift et al. (1991) have recently reviewed a number of
widely used models of cyclones, i.e., those due to Leith and Licht (1972), Muschelknautz (1970),
Dietz (1981) and Mothes and Lffler (1988). It emerges from their analysis that the models of
Dietz (1981) and Mothes and Lffler (1988) provide the best agreement with a very wide range
of experimental data reported in the literature.
Electrostatic Separators cover devices in which the main separating effect is migration of
electrically charged particles in an imposed electric field. These devices have a high collection
efficiency for a wide range of particle sizes, and are particularly suitable for submicrometer
solids. They have at the same time a high throughput and low pressure drop, which are the main
attributes for their wide use in the power generation industry. Operation under extreme
conditions of temperature (i.e., above about 800C) is difficult unless it is accompanied by high
pressures, as it will be described below. Apart from the problems of integrity of materials of
containment and construction, further considerations limit operation at elevated temperature and
pressure. Insulating materials are required to prevent excessive current leakage between the
electrodes. Breakdown of the gas between the electrodes must also be avoided. The minimum
potential gradient causing ionic breakdown of a gas generally decreases with increasing
temperature but increases with pressure. Therefore, at elevated temperature, there is generally a
minimum pressure, dependent on gas composition, below which a precipitator cannot be
operated. The electric properties of the particulate are also relevant: if it is highly resistive, then
collected dust retains its charge and reduces the efficiency of collection of further particulates. It
is also necessary for the collected dust to be removed from the precipitating electrode. This is
normally achieved intermittently by some mechanical action such as "rapping" and the detached
particles are allowed to settle to the base of the equipment. For this to be possible, the dust must
detach as agglomerates with high terminal settling velocity; thus some cohesiveness of the
collected dust is necessary to avoid excessive re-entrainment on electrode cleaning.
The process of electrostatic precipitation involves three stages:
a. charging of suspended particles by corona discharge;
b. migration and deposition of the charged particles under the influence of an applied
electrostatic field;
c. removal of the collected material from the collecting electrode and transfer to a suitable
receptacle outside the precipitator.
In industrial-scale electrostatic precipitation, stages (a) and (b) are generally combined and the
electrostatic field also produces the necessary corona discharge for particle charging. The
principles of operation are described by Strauss (1975) and by Bhm (1982). It is important to
establish a stable corona having adequate current to provide complete charging of the particles,
without getting sparkovers. The onsets of corona voltage and sparkover voltage depend on the
geometry of the electrodes, type of gas and the operating temperature and pressure. The corona-
starting voltage increases slowly as the pressure is increased, for both positive and negative
corona. The effect of pressure on sparkover voltage is more prominent. The sparkover voltage
increases rapidly as the pressure is increased above atmospheric, but reaches a maximum and
declines, to coincide eventually with the corona-starting voltage, see Robinson (1971). The
pressure at which the corona-starting and sparkover voltages coincide is called the "critical
pressure". The interaction of two opposing effects is considered responsible for the existence of
the critical pressure point as the pressure is raised (at constant temperature).
a. shorter mean-free paths impede ionization by collision, and so tend to raise the sparkover
level;
b. enhanced photoionization and reduced ion diffusion tend to facilitate streamer propagation
from the anode across the gas.
As the pressure increases, the initially dominant first effect gives way to the second, the streamer
develops across the gas, and at the critical pressure, spark breakdown ensues.
The effect of temperature has been studied by various authors and the results of the earlier work
have been summarized by Robinson (1971). Briefly, the earlier works indicated that, for
temperatures up to about 1073 K (800C), the current-voltage relations for positive corona were
a function of relative gas density only, but for negative corona the density and temperature had
independent influences for temperatures above 823 K (550C). At temperatures higher than
800C, thermal ionization was considered to become high enough to play a significant role and
eventually lead to breakdown. In this case a positive corona was predicted to be required for
most types of gases, as the space charge produced by thermal ionization would generally be
positive due to sweeping out of highly mobile electrons.
In more recent years, Bush et al. (1977 and 1979) investigated the range of temperature and
pressure in which stable corona discharge may be obtained, and established current-voltage
characteristics in a particle-free electrostatic precipitator for dry air, a simulated combustion gas
and a substitute fuel gas at temperatures up to 1366K (1093C) and pressures up to 35.5 bar.
Their results show that excessive current and sparkover due to thermal ionization, anticipated
previously, are not encountered within the range tested. The results also reveal the tendency for
the positive sparkover voltage to exceed that of the negative sparkover for temperatures above
533 K and low relative air densities (below about 2), a trend which had been observed by earlier
workers. However, experience with dust-laden gases has consistently shown that positive
coronas are inherently less stable than negative, and also yield a lower collection efficiency due
to lower corona currents.
The results of Bush et al. show that the critical pressure increases with temperature. The critical
pressure for negative corona discharge is much higher than for positive. It is worth noting that,
due to the increase of critical pressures with temperature, the range of pressure for stable corona
becomes wider at high temperatures. Thus precipitation could be made more efficient by
applying higher voltages as temperature and pressure increase together. However, these
conclusions are based on dust-free gases, and the performance evaluations reported so far cannot
confirm with confidence their extension to dust-laden gases.
The bulk of the particles passing through the charging zone acquire electric charges of the same
polarity as the discharge electrode; i.e., negative for negative corona. Two distinct mechanisms
are involved in the charging process:
a. Field charging, caused by bombardment of the particles by ions migrating under the influence
of the overall electric field.
b. Diffusion charging, resulting from attachment to the particles of ions which contact them in
the course of their random movement through the gas.
Field charging is the dominating mechanism for large particles, typically greater than 1.0 m. Its
rate depends on the field intensity as well as on ion concentration. Diffusion charging is
dominant for very fine particles, smaller than 0.1 m. In industrial precipitators the particle
residence time in the charging zone is large, and the particles reach their saturation charge level,
which may be estimated from Equation (19) due to Cochet (1956 and 1961). This equation
includes the effect of diffusion charging in the field charging process, hence taking into account
both mechanisms
(19)

where q is the charge acquired by a particle of diameter d
p
in the field of strength E
0
, k
p
is the
relative dielectric constant of the particle,
0
is the permittivity of free space, and is the gas
mean free path.
Under the influence of the electric field, charged particles migrate towards the collecting
electrode. The migration velocity, , is commonly calculated from the quasi-steady force
balance between Stokes' drag and the applied electrostatic force:
(20)

where E
p
is the precipitating electric field. Hence,
(21)

where C is the slip correction factor. In practice, is determined empirically because of various
complicating factors such as the presence of particle size and shape distributions, turbulence and
ionic wind.
The collection efficiency of an electrostatic precipitator is related to the migration velocity by the
celebrated Deutsch (1922) model. Details of this model and other technological issues such as
the grade efficiency and particle collection and discharge are given in the section on the
electrostatic precipitators.
Filters cover all devices in which the gas to be cleaned passes through a porous or permeable
medium which collects and retains particles carried by the gas. In a membrane-type filler, the
passages through the medium are comparable to or smaller than the particulate. Filtration then
occurs by mechanical obstruction, and the particulate builds up as a filter "cake" on the upstream
face of the filter. The filtration efficiency is essentially absolute, and virtually all the dust is
retained on the upstream surface of the filter. In other types of filter, such as granular
bed or fibre filters, the passages through the medium are typically large compared to the
particulate to be collected. The particulate is therefore collected initially within the bed by the
individual granules or fibres. This process is called depth filtration, and it continues until the
passages on the upstream side of the filter narrow down, leading to bridging over the openings
where a cake starts to form. In the initial stage of filtration here, the efficiency of the filter mid
retention of the particulate by the medium is of concern, while in the later stage of filtration,
where a cake has formed, the efficiency is high and the increase in the pressure drop across the
filter is of concern as in the membrane-type filters.
Filters require regular cleaning as the level of dust builds up on the filter. Various methods of
cleaning are used depending on the filter type. Membrane filters are sometimes used just once
and are then replaced or they are cleaned in situ by some intermittent mechanical process such as
vibration or reverse pressure pulse or reverse gas flow rate. A similar cleaning process is applied
to those filters which may operate initially by depth filtration, but rely on cake formation as the
main filtration mechanism. The granular type filters, on the other hand, employ beds of
unbounded filter elements, which are. usually removed continuously or intermittently for
regeneration of the filter medium. In some cases, the elements are cleaned in situ, as for a
membrane filter.
The range of materials used as filter medium is very wide indeed; it includes natural and man-
made fibres, and porous sintered polymer, metal and ceramic sheets and tubes. The main filter
medium requirements are:
a. durability of the medium at process conditions;
b. the ability of the dust to form a cake on the medium;
c. filter performance during the initial stage of cake formation following cleaning;
d. ability of the medium to withstand mechanical stresses during cleaning;
e. form of cake detached during cleaning;
f. "blinding" of the filter medium either by depth filtration or by permanent adhesion of cake to
medium.
Appropriate choice of filter medium depends on the process conditions and type of dust material.
High temperature applications impose a great constraint on the choice of filter medium. Recent
developments in this field have been addressed in a symposium on gas cleaning at high
temperatures [see Clift and Seville (1993)].
The process of filtration is analyzed in terms of three fundamental aspects of primary collection
of dust from the gas, retention or rebound of dust from the collector, and the effect of collected
dust on the filter structure.
For primary collection, the important mechanisms in general are: (i) deposition by diffusion; (ii)
gravitational settling; (iii) inertial deposition; (iv) direct interception; (v) electrostatic deposition.
Magnetic and thermal deposition may also occur depending on the application. The collection
efficiency of each individual mechanism is conventionally presented in terms of single fibre or
granule collection efficiency, E, defined by
(22)

with E a function of dust and granule or fibre size and charge, dust density, gas properties, and
bed voidage. For a particular dust size in a given filter, the penetration is defined as:
(23)

The single particle collection efficiency depends on type and structure of filter medium.
Furthermore, its relationship with the penetration is established through mathematical modelling
of filtration process. For granular type filters, the single particle collection efficiency of the first
four mechanical collection mechanisms has been summarized by Ghadiri et al. (1993). The
electrical deposition mechanism is given by Coury (1983), and the magnetic collection by Birss
and Parker (1981). For depth filters using fabric medium, various correlations for E have been
summarized by Strauss (1975) and Lffler (1971). For granular bed filters, the relationship
between E and f has been outlined by Ghadiri et al. (1993) and by Clift et al. (1981).
Applications to fluidised bed filters at elevated temperatures have been discussed by Ghadiri et al.
(1986). For fabric filters, the relationship between E and f has been analyzed by Schweers and
Lffler (1994), taking account of the distribution of inter-fibre spacings.
When a dust particle makes contact with a filter element, it is essential that it is retained rather
than rebounding, as it may lead to the re-entrainment of the particle. Retention is dominated by
short-range adhesive forces. The analysis of retention was initiated by Dahneke (1971), and
developed further by Stenhouse and Freshwater (1976), Hiller and Lffler (1978), Clift (1983)
and more recently by Ning (1995). The approach is based on an energy balance between the
kinetic energy of the approaching particle, E
i
, and the detachment energy, E
d
In addition to E
i
,
the particle acquires further energy in approaching the collector due to the very short van der
Waals forces, E
v
The energy at the instant of rebound is then e
2
(E
i
+ E
v
), where e is the
coefficient of restitution; the kinetic energy lost is due to the dispersion of elastic waves as well
as plastic deformation of the particle or the collector. In order to detach from the surface, the
particle energy must exceed the detachment energy, E
d
Thus the particle adheres if
(24)

The above inequality indicates a critical impact velocity above which a particle will rebound
from the filter element. The detachment energy Ed is the sum of energies due to van der Waals
forces, and electrostatic and capillary attractions if they are present. Plastic deformation enhances
cohesion because of the larger contact area than otherwise possible with elastic deformation
alone, and this has recently been modelled by Ning (1995).
Cake formation is essential in those filter types whose medium has large passages and the initial
stage of filtration is by depth filtration. Interparticle cohesion is essential for cake formation to
occur. The process starts with particle chain formation, leading to arching over the openings of
the filter, hence providing the foundation for build up of a cake. If the interparticle cohesion is
low or the span of the passages is too wide, the arch may be weak and it may fail under stresses
due to the pressure drop in the cake. In this case, the dust particles will break into the filter, i.e.,
pinhole formation, which causes long-term "blinding" of the filter. Ghadiri et al. (1989) showed
that pinhole formation can be reduced by electrical enhancement of the interparticle cohesion.
Coury (1983) has shown that particles that carry substantial electrical charges of both signs, such
as fly-ash, readily form a cake, where uncharged and uncohesive particles, such as coal char, fail
to form a cake under otherwise identical conditions. Cake formation is therefore strongly
influenced by interparticle forces including van der Waals, electrostatic and capillary forces.
Filter cleaning is another important aspect of filter operation. It should be carried out in such a
way that the cake is detached completely from the filter, i.e., without leaving patches of particles
still adhering to the filter. At the same time, the cake coherence should be preserved so that the
particles are not redispersed in the gas stream. Current practice in filter cleaning is largely
empirical. However, much of recent work has centered on developing an understanding of the
mechanisms involved from which some operational guidelines have emerged [see Cliff and
Seville (1993)].

S-ar putea să vă placă și