Vous êtes sur la page 1sur 8

1

Autoclave Design and Scale-up from Batch Test Data: A Review of Sizing Methods and Their Bases
Khosrow Nikkhah, H.A. Simons Ltd., Mining Group, 111 Dunsmuir Street, Suite 400, Vancouver V6B 5W3, BC, CANADA

ABSTRACT
A comprehensive understanding of the effects of autoclave
configuration (batch, semi-batch, continuous) on the residence
time necessary to achieve a target conversion is a prerequisite
for design of commercial scale autoclaves. Particle size
distribution, mineral and reagent concentrations, temperature
and mixing also influence the conversion.
This paper outlines the bases of using test data to predict
performance of commercial scale autoclaves. Recommendations
for coordinating test program design with selection of scale-up
method are provided. Examples are given to illustrate the
different levels of engineering effort needed to meet
corresponding levels of design optimization and accuracy.
INTRODUCTION
Commercial pressure leaching autoclaves have been in
operation since mid 1950s. Historical production data suggest
that these early autoclaves were originally designed with excess
capacity (Berezowsky, Collins, Kerfoot, and Torres, 1991).
Process development initially consisted of tests in batch
autoclaves. While in some cases this stage had been followed by
continuous pilot plant tests in multi-compartment autoclaves, in
many cases the batch test residence times were extended to the
continuous commercial stage using some factor, resulting in an
over estimation of the required autoclave volume. In some other
instances the continuous stage of testing had not been performed
or its results could not be interpreted as a continuation of batch
tests due to differences in ore composition, grind size and
reactor conditions. In such cases, it would be important to be
able to scale-up the batch test data to a multi-compartment
continuous mode using a sound approach that takes into account
both the physical and chemical aspects of the leaching reaction
as well as the reactor configuration.
Today, the fraction of capital cost of a hydrometallurgical plant
associated with the pressure autoclaves could be as high as 30%
of the total capital expenditure. Moderate over-estimation of
volume requirements can significantly drive-up such costs,
either through buying larger autoclaves or (since transporting
such units can be difficult) unneeded extra units. This can be
avoided by a thorough understanding of the effects of autoclave
configurations on scale-up to the commercial-scale units. While
it is important to avoid over-design and its associated costs, the
engineer must also ensure that possible variations in ore grade
and quantities will not lead to shortfalls in extraction efficiency
for the design residence time.
FACTORS AFFECTING AUTOCLAVE SIZE
In a recent paper on design of autoclaves for pressure leaching
of nickel laterites (King, 1996), the following have been listed
as the determining factors in the pressure autoclave size:

solids throughput and size distribution


selected autoclave circuit
slurry density
retention time

temperature
number of operating trains
shipping restrictions.

Retention time is independent of throughput as long as all of


the other aforementioned factors are kept constant. Assuming
that reaction condition and feed properties are uniform, several
approaches have been followed to arrive at scale-up factors to
predict continuous retention times from batch laboratory tests.
While most of these methods are based on assumptions that
include the shrinking of solids particles that are dissolving by
reaction with an aqueous reagent, the earliest method considered
was an extension of homogeneous scale-up calculations.

SCALE-UP ASSUMING HOMOGENOUS SYSTEM


The simplest approach for sizing of autoclaves was that used
by OKane (1963) in his paper on Pressure Leach Autoclave
Design. The method used was based on an earlier work
(Caddell and Hurt, 1951). It consisted of taking a leach curve
(mineral conversion versus time) from the batch test results and
using it graphically to determine each continuous autoclave
compartment residence time. Hence the overall autoclave size
for a given target conversion could be calculated. This method
assumes that the solubilization reaction is so fast that the
heterogeneous leaching reaction approaches a homogeneous
one, and the reaction rate can be represented merely in the units
of time-1.
In developing their method, the authors assumed that the
composition of the reacting species is constant within the
autoclave compartment. In a laboratory batch test, however the
reaction rate is initially high due to the high reactant
concentration. They proceeded to assume that in an autoclave
compartment, the reaction rate is equal to the minimum that
prevails at the end of a corresponding batch cycle resulting in
the same concentration and can be represented by the slope of
the leach curve at that point. The graphical method comprises
drawing a tangent to the leach curve at the designated exit
composition point for the final compartment. This is done by
starting at the end compartment and extending the tangent until
an assumed residence time was reached. The point on the curve
horizontally across from where the tangent ends, is where one
should continue the process by drawing the tangent for the
penultimate compartment unit and so on.
Figure 1 shows this approach for a four-compartment
autoclave, where a 95% conversion has to be achieved. If after
repeating this procedure four times for a pre-assumed
compartment residence time, the horizontal projection of the
tangent drawn for compartment one does not intersect the origin,
then the process would have to be repeated, with a new
assumption for the unit residence time.

2
reactions in hydrometallurgy are prime candidates for analytical
treatment in this manner.
Assuming that minerals are in the form of spherical particles
that react in an aqueous phase, this treatment can be applied to
pressure leaching operations that operate under batch and
continuous configurations. In this way a more accurate model
of a leaching operation can be developed.
For both homogeneous and heterogeneous systems reaction
rate can be described to be a function of both composition and
temperature. In heterogeneous systems concentration can be
represented by the size of the spherical solid particle. For a
component A(q) in the liquid phase reacting with a solid B(s) by
the chemical equation:
Figure 1. Graphical Calculation of Reaction per Stage,
(OKane, 1963).

In addition to obvious short comings of using a method


intended for liquid-liquid reactions for mineral solubilization, to
apply this method graphically would render a rather subjective
result. It is possible to use spreadsheets and apply computer
programming to this method so as to arrive at an accurate
implementation of the method. In such an approach, the
engineer faces the challenge of developing a true quantitative
representation of the leach curve and the corresponding tangents
(reaction rates).

A(q) + bB (s) products

(1)

the quantity of the reactant is proportional to the available


surface of the unreacted particle core. This assumption is based
on the fact that physical diffusion processes that transfer
reactants and products to the solid-liquid reaction sites are fast
enough for the surface reaction to be rate controlling, i.e.,

b dN A
1 dN B
=
= bk s C Aq
2
4rc dt
4rc2 dt

(2)

SCALE-UP METHODS FOR HETEROGENEOUS


PARTICLE-FLUID SYSTEMS
The shrinking core model described by Lievenspiel (1972) for
solid-fluid systems forms the basis for the quantitative
comparison of mineral leaching operations in batch, plug flow
and mixed-reactor modes. The complexity of such models can
increase to accommodate some of the other factors that affect the
reaction kinetics. In essence they range from simplified scale-up
calculations to full fledged computer simulation techniques that
require extensive engineering and computer programming input.
All methods rely on availability of kinetic data from batch tests.

where NB is the moles of reactant B present, rc is the unreacted


core diameter, ks is the first order rate constant for the surface
reaction and CAq is the concentration of the aqueous reactant at
the particle surface. Writing NA in terms of the shrinking sphere:

dr
1
2 drc
= b c = bk s C Aq
2 b 4rc
dt
dt
4rc

(3)

Theoretical Background of the Shrinking Core Model


Leaching of ores and concentrates involves either a gaseous
reactant such as oxygen or an aqueous one in the form of acid or
caustic solutions. Whether the reactant is a gas or liquid, it has
to be in the aqueous phase prior to reacting with the particular
solid component of the feed. Therefore, in solid-liquid reactions
usually more than one process is involved in the overall
solubilization of the mineral. In addition to the chemical
reaction taking place at the solid liquid interface, there is a
physical process involving the diffusion of the gaseous reactants
into the liquid phase and through the solid pore in the aqueous
phase. Octave Lievenspiels (1972) original treatment of
equations for heterogeneous reactions extends to derivation of
reaction rates for shrinking spherical solid particles under both
chemical reaction and diffusion control regimes. These
equations can be applied to both batch and continuous
configurations to obtain the appropriate expressions for
residence time in terms of reactant conversion. Leaching

where b is the molar density of reactant B. On integration, time


for complete conversion of a particle with original radius R, is
represented by:

b R
bk s C Aq

(4)

As the time required for the increase in fractional conversion


according to equation 4 is shown to be proportional to reduction
of the particle radius, the dimensionless time for solubilization
of a spherical particle can be represented as:

= 1

rc
= 1 (1 X B )1/ 3
R

(5)

In a similar fashion for a diffusion control regime the


dimensionless time for conversion of the mineral according to
the shrinking core model, is given by:
2

2/ 3
r
= 1 c = 1 (1 X B )

(6)

where R is the original radius of the spherical particle and XB is


the fractional conversion of the mineral.
Equations 5 and 6 represent the bases of several different
approaches to modeling of mineral solubilization reactions.
These techniques have had varying degrees of success
depending on the accuracy of model assumptions.
Review of Scale-up Methods Based on the Shrinking Core
Model
Henein and Biegler (1988) using the bases of shrinking core
model, first order reaction rate and ideal, continuous stirred tank
(CSTR) and plug flow reactors (PFR) considered three different
cases: a single mixed tank, n mixed tanks and a single plug flow
reactor. In each case they compared the reactor volumes for both
surface and diffusion controlled mechanisms for different
conversions. Although the assumption that the size distribution
of the solid feed is monosized, makes this model less rigorous,
the design equations are simpler and provide a rapid tool for
calculating reactor size, based on batch test data.
Henein and Biegler (1988) derived the dimensionless
Damkohler number, Da (residence time) for both plug flow and
individual and a series of mixed reactors based on surface
reaction controlling:
(Das)PFR = 3 (1 - (1-)1/3)

(7)

(Das)CSTR = /(1-)2/3

(8)

(Das)CSTRj = (j-j-1) /(1-j)2/3

(9)

where is the fraction of the solid reacted.


Knowing the batch residence time from test work, one can
calculate a theoretical reactor volume scale-up factor for going
from batch and or plug flow to a single or n ideal mixed
reactors.
Results obtained using this model (Henein and Biegler, 1988)
have been shown to slightly over estimate the scale-up factors
because they do not account for reduction in particle size and
the changing particle size distribution with conversion at the
outlet of succeeding compartments (Sepulveda and Herbst,
1978; Crundwell, and Bryson, 1992). However, if both the
batch tests and the continuous campaigns were to be performed
with a feed of the same size distribution, it can provide a fairly
accurate assessment of scale-up factor required for the initial
attempt in sizing a hydrometallurgical reactor.

Pritzker (1993) extended the analysis of Henein and Biegler


(1988) by considering the simultaneous depletion of fluid phase
and solid phase reactants. This approach allowed the author to
develop design equations, which include feed composition and
mode of operation. As an example, it was shown that equations
could be applied to hydrometallurgical leaching systems and
determine residence times and inter-stage conversions for cocurrent and counter current multi compartment systems. Pritzker
(1993) derived a series of equations that can be applied to both
batch tests and CSTRs under pore diffusion, interfacial kinetics
and mixed reaction control modes. Although he used the same
approach as Henein and Biegler (1988) by not considering size
distribution of the solid feed, he was able to improve their model
for a situation, where the depletion of the fluid phase component
has to be considered.
In recent years there has been more emphasis on the effect of
particle size distribution as well as the reactor heat and mass
balance on the overall reactor size and conversion. Crundwell
and Bryson (1992) have shown that a greater degree of accuracy
in prediction of scale-up factors for autoclave-sizing can be
obtained by forgoing the monosized assumption of Henein and
Biegler (1988) and Pritzker (1993). Assuming that the batch test
size distribution is the same as that of the continuous test, and
then assuming a mean particle size, is fairly accurate for the first
reactor compartment. Deviation from the assumed size
distribution occurs only in the subsequent compartments as
there is a unique and different particle size distribution leaving
the first reactor, even if the feed was truly monosized. The
model would therefore not be accurate unless it included the size
distribution of the feed to the subsequent reactors.
Bartlett (1971) presented one of the earlier adaptations of the
shrinking core model coupled with the effect of size distribution
based on the segregated flow model for metallurgical reactors.
The segregated flow model is based on the assumption that the
ore slurry is composed of small elements of the fluid-solid
mixture, which behave as moving batch reactors. For the case of
a well agitated autoclave unit, where the mineral particles can be
represented as shrinking spheres controlled by surface reaction,
Bartlett (1971) defined the yield, X, by the following equation:

1 (1 X )1/ 3

M
( rs )t
c
=
Ri

(10)

where Ri is the radius of the ith particle, (M/c) is the molar


volume of the solid particle and -rs can be equated to the
chemical reaction rate for the solubilization of the mineral.
Assuming that the rate of solubilization of the mineral is equal
to the rate of shrinking of the whole particle, then:

M
1
1 (1 X )1/3 = ( kc C An )t
Ri
c

(11)

For this case Bartlett (1971) derived the following expression


for the unreacted fraction of the feed in a batch reactor based on
the shrinking core model:

1 X = 1

(12)

where is the time for complete conversion at constant flux of


reactant to the particle surface and can be represented by:

R
1 X = 1 i
( Ri )

(Ri) =

, and

Ri

(13)

( M / c )( rs )

For continuously stirred tank reactors, Bartlett (1971) based his


treatment on the concept of non-ideal flow in the form of the
segregated flow model, where E(t), the age distribution of the
particles in a single mixed reactor is given by:

1
E ( t ) = e t / t
t

(14)

(15)

Later in this paper, a discussion is presented on the predictions


for the performance of a continuous leaching operation based on
both the graphical method by Caddell and Hurt (1951), the
monosized method of Henein and Biegler (1988) and the
potential extra accuracy obtained by assuming a particle size
distribution in the solid feed. Continuous reactor scale-up
factors for high pressure oxidation of refractory gold ore based
on batch leach curves at the McLaughlin Mine reported by
Turney, Smith, and Janhunen (1988) are calculated and
compared to the available continuous data. Effect of particle
size distribution is also considered.

where t is the particle residence time and t is the average


residence time for a backmix reactor. Hence, the particle
residence time distribution coming out of the jth reactor of j
reactors in series, could be given by:

E (t ) =

j (t / t )
_

j 1

In both of the above studies, the treatments provided for


modeling of a cascade of reactors based on non-ideal flow
(segregated flow Model) gain in accuracy over monosized
particle model. However, both Sepulveda and Herbst (1978)
and Crundwell, and Bryson (1992) developed this approach
further, by using the population balance model. This is a
number balance that can be used for the description of a
particulate phase in slurries. Hulubert and Katz (1964) gave an
original introduction of the concept of population balance. In
this way, using differential equations, one can represent the
particle size distribution based on an assumed rate of change of
particle size.
In addition to consideration of particle size distribution, there
is another effect that is unique to a series of mixed reactors: the
principle of non-ideal flow shows that for particles that
experienced short residence time in the first reactor, it is not
likely that they exhibit the same lower residence time in all of
the subsequent compartments.

the diffusivity of the reacting components in the liquid such as


oxygen through a solid ash layer. The derivation of the
expression for the unreacted fraction of the mineral in this case
also included exit age distribution function E(t) and the
segregated flow model.

j (t /t )

t ( j 1)!

Bartlett (1971) then proceeded to include both the expression


for E(t) as well as the values of Gates-Gaudin-Schuhmann
(GGS) parameter in the expression for the fraction unreacted in
the case of several CSTRs in series. The results were plotted for
a series of GGS grinding parameters for single stage and multi
stage CSTR reactors. These results were compared with similar
results for batch reactor tests and are presented in form of the
typical curves of fraction of feed unreacted versus dimensionless
time for both batch reactors and mixed flow reactors. Using
these curves one can determine the effect of going from
laboratory batch tests to continuous operation and choose the
scale-up factor for residence time accordingly. The main
advantage of Bartletts work (1971) was that the effects of
particle size distribution on the scale-up factors from batch to
mixed reactors with different number of compartments could be
evaluated. As expected, reduction in particle size leads to a
shorter retention time for achieving a target conversion.
A later work by Reuther (1979) has also adapted the shrinking
core model for determining residence time for a cascade of
stirred tank reactors. In this model which was applicable to
reactions involving sulfur removal from coal, only the treatment
for the case of diffusion controlled reactions was presented. In
this model it was necessary to have good experimental data on

Other attempts at scale-up, modeling and eventually simulation


of high pressure multi-compartment autoclaves have used
population balance model and the effects of heat and mass
transfer on the autoclave design. These efforts, although
increasing the accuracy of the exercise, need extensive
engineering and computing efforts. They are of special value for
process control and operator training, if they can be extended to
dynamic modeling of the process.

Sample Test Data


The sample test data used as an example in this paper is that
reported by Turney, Smith, and Janhunen (1988) on the
McLaughlin refractory gold ore. The test work reported
consisted of both batch and continuous pilot scale. The batch
test is not described in any detail. It is assumed that the batch
tests were performed in Parr batch autoclaves with the same feed
and reactor conditions as the reported continuous tests. The
continuous tests were conducted in a four compartment
laboratory autoclave at a feed rate of 6.6 lb./hr with a retention
time of about 1 hours at 160 and 175 oC and pressure of 175
psig. The oxygen over-pressure was reported to be 30 psi.
Each compartment was equipped with a thermowell, an oxygen
sparge tube and a sample dip tube. Adequate agitation was
achieved by axial flow impellers and baffles. The useful

5
volume of each compartment was slightly short of 2 gallons.
Heating was provided to each compartment by a direct gas
burner.
Figures 2 and 3 show the batch and continuous leach curves
for the pressure oxidation tests. Figure 2 illustrates the rate of
sulfur oxidation for the batch test at 165 and 175 oC. It was
shown that in a similar fashion to first order shrinking core
reactions under surface reaction control, the reaction rate while
increasing with temperature, decreases with % sulfur converted.

90
80
160 C
175 C

60
50

100

40
90

30

80

20

70

10
0
0

20

40

60

80

100

120

% S Oxidation

% S Oxidation

Assuming that these tests were performed under such adequate


level of agitation that surface reaction was the rate controlling
mechanism, one can apply equations 7 and 9 derived by Henein
and Biegler (1988) and arrive at the residence times required
under continuous operation. These theoretically predicted
results can then be compared with the actual continuous results.
Figures 4 and 5 illustrate such a comparison for the McLaughlin
test data.
Because the model of Henein and Biegler, (1988) assumes
monosized particles, the mean reaction rates in the latter
compartments are overestimated. In order to match this model
to the target conversion at the exit of the final compartment, the
conversion for the first compartment is calculated to be lower
than the actual result obtained in the continuous tests.

100

70

SCALE-UP OF TEST DATA

60
Continuous Test
50

Model

40
30

Oxidation Time, minutes


20

Figure 2. Sulfur Oxidation in a Batch Autoclave for


McLaughlin Refractory Gold Ore, (Turney, Smith, and
Janhunen , 1988).
100.00

80.00
160 C
175 C

20

40

60

80

100

120

Oxidation Time, Miniutes

50.00
100

40.00
90

30.00

80

20.00

70

10.00
0.00
0

20

40

60

80

100

120

Oxidation Time, minutes

% S Oxidation

% S Oxidation

60.00

Figure 4. Comparison of Prediction of Conversions and


Residence time scale-up factor (Model: 1.246; Testwork:
1.333) for a Four Compartment autoclave based on Henein
and Bieglers Monosized Model (1988) and the Continuous
Leaching Results for McLaughlin Refractory Gold Ore at
160 oC, (Turney, Smith, and Janhunen , 1988).

90.00

70.00

10

60

Continuous Test

50

Model

40
30
20

Figure 3. Sulfur Oxidation in Four Compartment


Continuous Autoclave for McLaughlin Refractory Ore,
(Turney, Smith, and Janhunen , 1988).

10
0
0

20

40

60

80

100

120

140

Oxidation Time, Minutes

Figure 3 shows the same trends for changes in reaction rate


with temperature and %sulfur converted. However, in this case,
where the testing was continuous, more residence time is
required to achieve a target conversion. The main focus of the
current paper is the accurate prediction of this scale-up in
residence time.

Figure 5. Comparison of Prediction of Conversions and


Residence time scale-up factor (Model: 1.376; Testwork:
1.333) for a Four Compartment autoclave based on Henein
and Bieglers Monosized Model (1988) and the Continuous
Leaching Results for McLaughlin Refractory Gold Ore at
175 oC, (Turney, Smith, and Janhunen , 1988).

6
1
0.98
0.96

Conversion

However, it would be safe to use this method for the


estimation of the total autoclave size at the preliminary stages of
an engineering study as a quick design method. In fact only the
ultimate target conversion from the batch test data is needed to
arrive at a fairly good estimate for the autoclave size. The
factors for scale-up of volume from batch to continuous
operation based on the method of Henein and Biegler (1988)
were calculated to be 1.246 and 1.376 for leach tests at 160 oC
and 175 oC respectively. The pilot plant data for the
McLaughlin ore suggested a scale-up factor of 1.333 in both
cases.

0.94
0.92
0.9

Model Compartment 1
Test Compartment 1
Test Compartment 2
Model Compartment 2
Test Compartment 3
Model Compartment 3
Test Compartment 4
Model Compartment 4

0.88
0.86
0.84
0.82
0.75

0.8

0.85

0.9

0.95

1.05

1.1

1.15

Slurry flowrate, m /min

In addition, application of the graphical method outlined by


OKane (1963) for the aforementioned McLaughlin leach curves
resulted in scale-up factors of about 2.00 and 2.66 for the tests at
160 oC and 175 oC respectively. This approach considerably
over-estimates the required residence time and the results are
therefore of value, only when a very rough and conservative
prediction of the residence time is needed.

Figure 6. Comparison of Model and Plant Data for Two


Different Slurry Flow Rates for the Trail Pressure Leach
Autoclave (Crundwell and Bryson, 1992).

Consequently, incorporation of the shrinking core model


coupled with the grind size distribution of feed in to the first
compartment and modeling of the particle size distribution in
the subsequent compartments is necessary in the latter parts of
engineering design work. This is specially important in the case
of pressure oxidation of refractory gold ores, as size of the first
compartment has a direct bearing on the autoclave heat balance
and operation of the autoclave under the optimum conditions
(with possibility of having autothermal oxidation reactions in
the first compartment).

Application of these models would gain in accuracy with


availability of data that goes beyond what would be feasible
during initial laboratory test programs. Furthermore, they are of
great use both in optimization and the detailed design stage,
where it is necessary to analyze the process effectiveness for
variation in ore or concentrate grind size, slurry density, steam,
cooling water and recycling requirements. Extension to dynamic
modeling would be the logical next step, where computer
simulation in addition to design of the autoclave could be used
for investigation of control philosophy, HAZOP studies and
operator training. Rubisov and Papengelakis (1995) have
published a recent work in this area, which attempts to address
some of these objectives.

The studies by Sepulveda and Herbst (1979) and Crundwell


and Bryson (1992) provide a sound basis for the development of
more accurate models that include particle size distribution.
Sepulveda and Herbst (1979) used the population balance model
to describe multistage leaching systems. They showed that their
models prediction compared fairly well with experimental data
for the acid ferric sulfate leaching of chalcopyrite.

Review of computer simulation studies of high pressure


autoclaves is beyond the scope of this paper; however
application of one or more of these models to a real autoclave
design and operation problem can form the basis of a very
interesting future study.

Crundwell and Bryson (1992) show the effect of product size


distribution on conversion in a single mixed reactor for the case
of reaction control by product diffusion. The model was also
compared with results from operation of zinc pressure leach
reactor in Trail, British Columbia, which was thought to be
operating under control of both surface reaction and product
layer diffusion. This comparison for the four compartment
autoclave at Trail is shown in Figure 6. The experimental results
provided for zinc pressure leach were at two different flow rates
and apparently this commercial unit could operate at up to 200%
above its design capacity!
More recent attempts at simulation of autoclave performance,
have taken into account particle size distribution and heat and
mass balance. Most of these studies have been based on
detailed research and development work, which is beyond a
scope of a single engineering project. They include a general
treatment for modeling of multistage continuous reactors at
steady state (Dixon, 1996), mathematical modeling of refractory
gold ores (Papengelakis, Berk and Demopoulos 1986, 90),
(Papengelakis and Demopoulos 1992, 1992, 1992, 1993) and
computer simulation of zinc pressure leach process (Baldwin
and Demopoulos, 1995), (Baldwin, Demopoulos and
Papengelakis, 1995).

BATCH TEST PROGRAMS


Accuracy of scale-up methods for sizing autoclaves are based
on the soundness of the laboratory test data available. Choosing
a correct approach for implementation of a batch test program is
therefore of fundamental importance in sizing a continuous
autoclave. Table 1 shows a typical program in the form of a
results summary for laboratory batch tests, where a gaseous
reagent such as oxygen is used to dissolve mineral constituents.
Prior to following such a program, it is necessary to verify that:
1.

Adequate agitation is provided, such that the reaction rate


is not dependent on mass transfer rate of the aqueous
reactant to the mineral surface. This is achieved by
performing batch leach tests for a fixed period of time
under constant reaction conditions and varying the degree
of agitation, until it no longer effects the rate of dissolution.

2.

Variables, such as reactor temperature and pressure, overpressure of the gaseous reactant, slurry density and grind
size are fixed either by relying on published data or
performing preliminary tests where one or more of these
variables are varied for a fixed residence time.

7
Having optimized the reaction conditions, the next step is to
perform a series of tests at varying residence times to produce a
leach curve showing the progression of mineral dissolution with
residence time. Performance of repeat tests to prove the
reproducibility of results is also important.
Other test results worth following are residual solids
constituents, other metal values and acid in solution. In
performing autoclave batch test programs, it is important to
maintain the focus on the goal of the entire process. For
example, when performing high pressure sulfuric acid leach tests
for a laterite, it is important to limit the acid addition to the point
where the quantity of nickel in solution reaches a maximum.
Similar high dissolution of nickel and cobalt may be achieved
by continuing to add acid, while larger quantities of other ore
constituents (e.g., iron and aluminum) are being dissolved. It is
required to follow the dissolution of these metals as well, as
high concentration in the leach solution would adversely affect
the usage of acid and neutralization reagents.
In leaching reactions, it is always necessary to consider the
complex chemistry of the process in its entirety, before deciding
on the range of results that is adequate for a batch test program.
For example, in the case of high pressure acid leach of a laterite,
while dissolution reactions are the focus, the precipitation
reaction of iron which results in the formation of hematite also
needs to be completed for a useful set of results to be obtained
for scale-up.
If the test program is performed in a batch mode, it can be
theoretically analyzed and related to a continuous operation, for
the purposes of design and scale-up. However, if the procedures
used in the laboratory tests are unnecessarily too complex, it
would be difficult to quantify the results and use them as the
basis of scale-up.
In a recent paper, (Gathje, Oberg and Simmons, 1995) a new
approach to laboratory testing for autoclave design and scale-up
is proposed, which suggests that the mode of testing should be
semi-continuous in order to be able to obtain a better
understanding of scale-up to the continuous operation. This
method starts with a normal batch test as described for the
McLaughlin sample. This test provides the initial sample for the
semi-continuous method. Then a portion of the slurry is
discharged from the batch test autoclave and replaced with fresh
feed. Each cycle is repeated until an equilibrium is reached.
The problem with this method is that it cannot be quantitatively
treated as either a batch or a continuous operation. It is
therefore difficult for it to be considered as the basis of scale-up,
unless it is used in conjunction with an unnecessarily complex
theoretical treatment which takes into account the addition and
withdrawals of mass into the reactor. This approach replaces
neither the batch nor the continuous pilot plant stage of testing.

Table 1. Typical Summary of Tests for Dissolution of Mixed


Sulfides from Laterite Processing in a Batch Autoclave.

Reactor
Temp.

Oxygen,
OverPressure
-

Ni, gpl

Co, gpl

Fe, gpl

Species

Solids,
Wt. %

Reactor
Pressure

Ni
Co
Fe
Mn

50
5
0.2
0.01

Slurry,
Wt. %:
Time,
minutes
30
45
60
75
90
105
120
120+
Repeat
Runs

10.5
Solids
Wt. %
-

CONCLUSIONS
Scale-up of leaching reactions have been performed using
models based on the shrinking core principle with and without
consideration of the particle size distribution. While models
based on the monosized particle distribution are of adequate
degree of accuracy for pre-feasibility and feasibility studies, it is
essential to consider particle size distribution in more detailed
engineering designs. Complete batch test data must be available
for the development of these models. Performance of laboratory
tests and the equipment used should be such that scale-up to
continuous operation could be achieved by a relatively simple
quantitative treatment.
Because of complexities of the reactions involved in pressure
leaching and characteristics that are particular to different modes
of operation, scale-up models cannot replace pilot plant
campaigns based on continuous operation.

8
REFERENCES
Baldwin, S.A. and Demopoulos, G.P., 1995, Assessment of
Alternative Iron Sources in the Pressure Leaching of Zinc
Concentrates Using A Reactor Model, Hydrometallurgy, 39,
pp. 147-162.
Baldwin, S.A., Demopoulos, G.P., and Papengelakis, V.G.,
1995, Mathematical Modeling of the Zinc Pressure Leach
Process, Metall. and Mat. Trans., B 26B, pp. 1035-1047.
Bartlett, R.W., 1971, Conversion and Extraction Efficiencies
for Ground Particles in Heterogeneous Process Reactors,
Metall. Trans., B 2B, pp.2999-3006.
Berezowsky, R. M. G. S., Collins, M.J., Kerfoot, D.G.E and
Torres, N., 1991, The Commercial Status of Autoclave
Technology, JOM, February, pp.9-15.
Caddel, J.R. and Hurt, D. M., 1951, Principles of Reactor
Design, Chemical Engineering Progress, , 47, 7, pp. 333-338.
Crundwell, F.K. and Bryson, A.W., 1992, The Modelling of
Particulate Leaching Reactors- the Population Balance
Approach, Hydrometallurgy, 29, pp. 275-295.
Dixon, D.,1996, The Multiple Convolution Integral: A New
Method for Modeling Multistage Continuous Leaching
Reactors, Chem. Eng. Sci., 51, 21, pp. 4579-4767.
Gathje, J., Oberg, K.C., and Simmons, G., 1995, Pressure
Oxidation Process Development: Beware of Laboratory Testing
Results!, Presentation at the SME Annual Meeting, Denver,
Colorado, March 6-9.
Henein H., and Biegler, L.T., 1988, Optimization of Reactor
Volumes for Topochemical Reactions,, Trans. Inst. Min.
Metall., 97, pp.C215-223.
Hulburt, H.M., and Katz, S., 1964, Some Problems in Particle
Technology: A Statistical Mechanical Formulation, Chem.
Eng. Sci., 19, pp. 574.
King, D. H., 1996, Autoclave Design for Pressure Leaching of
Laterites, Presentation at Nickel/Cobalt Pressure Leaching and
Hydrometallurgy Forum, Perth, Australia, May 13-14.
Lievenspiel, 1972, Chemical Reaction Engineering, 2nd
Edition, John Wiley and Sons.
OKane, P.T., 1963, Pressure Leaching Autoclave Design,
Presentation at the 13th Annual Canadian Chemical Engineering
Conference, Montreal, Quebec, October.

Papengelakis, V.G., Berk, D. and Demopoulos, G.P., 1986,


Modelling and Simulation of a Batch Reactor for An Oxygen
Pressure Leaching System, Presentation at Hydrometallurgical
Reactor Design and Kinetics, New Orleans, Louisiana, USA,
March 2-6.
Papengelakis, V.G., Berk, D. and Demopoulos, G.P., 1990,
Mathematical Modelling of an Exothermic Leaching Reaction
System: Pressure Oxidation of Wide Size Arsenopyrite
Particles, Metall. Trans., B 21B, pp. 827-837.
Papengelakis, V.G. and Demopoulos, G.P., 1992, Reactor
Models for a Series of CSTRs with a Gas-Liquid-Solid Leaching
System: Part I, Surface Reaction Control, Metall. Trans., B
23A, pp. 847-856.
Papengelakis, V.G. and Demopoulos, G.P., 1992, Reactor
Models for a Series of CSTRs with a Gas-Liquid-Solid Leaching
System: Part II, Gas Transfer Control, Metall. Trans., B 23A,
pp. 857-864.
Papengelakis, V.G. and Demopoulos, G.P., 1992, Reactor
Models for a Series of CSTRs with a Gas-Liquid-Solid Leaching
System: Part III, Model Application, Metall. Trans., B 23A,
pp. 865-877.
Papengelakis, V.G. and Demopoulos, G.P., 1992, On the
Attainment of Stable Autothermal Operation in Continuous
Pressure Leaching Reactors, Hydrometallurgy, 29, pp. 297318.
Pritzker, M.D., 1993, Reactor Design for Topochemical
Leaching Reactions Taking into Account Simultaneous
Depletion of Solid and Aqueous Reactants, Trans. Inst. Min.
Metall., 102, pp. C149-C158.
Reuther, J.A., 1979, Reaction in a Cascade of Continuous
Stirred Tank Reactors of Particles Following the Shrinking Core
Model, Can. J. Chem. Eng., 57, pp. 242-245.
Rubisov, D. H. and Papengelakis, V.G., 1995, Model -Based
Analysis of Pressure Oxidation Autoclave Behaviour During
Process Upsets, Hydrometallurgy, 39, pp.377-389.
Sepulveda, J.E. and Herbst, J.A., 1978, A Population Balance
Approach to the Modeling of Multistage Continuous Leaching
Systems, 1979, AIChE Symp. Ser., 57, pp. 41-55.
Turney, J.R., Smith, R.J. and Janhunen, W.J. Jr., 1988, The
Application of Acid Pressure Oxidation to the McLaughlin
Refractory Ore, Precious Metals 89, Edited by Jha, M.C. and
Hill, S.D., pp. 25-45.

Vous aimerez peut-être aussi