Você está na página 1de 9

Dynamic modelling of copper solvent extraction mixersettler units

C.M. Moreno
a
, J.R. Prez-Correa
a,
*
, A. Otero
b
a
Department of Chemical and Bioprocesses Engineering, Ponticia Universidad Catlica de Chile, Casilla 306, Santiago 22, Chile
b
Mining Centre, Ponticia Universidad Catlica de Chile, Casilla 306, Santiago 22, Chile
a r t i c l e i n f o
Article history:
Received 30 January 2009
Accepted 4 September 2009
Available online 7 October 2009
Keywords:
Extractive metallurgy
Hydrometallurgy
Simulation
a b s t r a c t
The copper Leaching, Solvent Extraction and Electrowinning circuit (LXSXEW) is one of the most effec-
tive processes for extracting copper from low grade ores. This work focuses on the liquidliquid extrac-
tion SX sub-process, since many difcult to handle operational problems within LXSXEW circuits are
related to SX malfunction. Controlling these problems better can reduce operational costs and increase
recovery and copper production. Realistic dynamic simulation is a standard tool nowadays to design
and assess more effective control strategies. In this work we present a general dynamic model for SX
mixersettlers, and applied it to two different units, one located in an extraction stage and the other
in a stripping stage of a copper plant. The model includes a non-trivial settler hydrodynamics represented
by a set of nonlinear differential equations for both mixer and settler units. The mixer is modelled as a
continuous stirred tank reactor and the settler as a hydrodynamic circuit combining series and parallel
connections of continuous stirred tank and plug ow reactors. The model was calibrated with industrial
plant data, resulting in realistic simulations of outlet copper concentrations. Using the proposed model,
we obtained better tting than that achieved with simpler settler models that include only a time delay.
The model tting parameters provide sufcient exibility to accurately reproduce the dynamics of differ-
ent units in industrial plants.
2009 Elsevier Ltd. All rights reserved.
1. Introduction
Due to increasing copper world demand, there is a strong incen-
tive to apply different processes to copper extraction from low
grade ores. The copper Leaching, Solvent Extraction and Electro-
winning circuit (LXSXEW) (see Fig. 1) is one of the most effective
processes for this purpose. In this work, we focus on the liquidli-
quid extraction SX process, since many of the difcult to handle
operational problems within this circuit are related to SX malfunc-
tion. Critical problems normally found in SX plants, which have a
considerable impact in extraction efciency and selectivity, are:
crud formation, organic and aqueous phase entrainments, and var-
iable and unpredictable phase separation times in the settlers
(Bergh and Yianatos, 2001).
For example, uncontrolled organic and aqueous ow-rates may
cause inefcient phase separation, leading to entrainment and pro-
cess instability. These, in turn, reduce cathodes quality, increase
the consumption of valuable chemicals, contaminate the electro-
lyte solutions, cause premature corrosion of expensive anodes
and increase organic phase losses. Additionally, copper concentra-
tion and pH are usually sampled and regulated manually by oper-
ators. Since the samples are analyzed in the laboratory, taking 4 h
in average, it is not surprising that process regulation is poor. As a
result, strong disturbances are unavoidable and many of them may
pass unnoticed and uncontrolled for at least 4 h. Hence, it can be
argued that many of the SX operational problems are related to
the lack of effective monitoring and control systems specic for
this process.
Consequently, on-line measurement of copper concentrations
in the main process streams together with an effective automatic
control strategy, should stabilize the process, improve cathodes
quality and homogeneity, and reduce operational costs. Unfortu-
nately, it is not possible to measure on-line copper concentrations
with affordable and reliable instrumentation. In addition, it is
expensive and risky to develop and test control strategies directly
in the process plant. Alternatively, dynamic simulation can be used
effectively to develop and test estimation and control strategies
before they are implemented on-line, reducing the risks of
catastrophic operational events, and the time and costs of
development.
Despite that several SX process models have been presented in
the literature, many of them are steady state or either too simple to
represent realistic industrial data or too complex to calibrate.
Moreover, there are still many phenomena in SX processes that
are not well understood. For example, Van Bochove et al. (2000)
developed a thermodynamic model that predicts exactly the iso-
therms of equilibrium. However, this model is very complex and
0892-6875/$ - see front matter 2009 Elsevier Ltd. All rights reserved.
doi:10.1016/j.mineng.2009.09.003
* Corresponding author. Tel.: +56 3544258; fax: +56 3545803.
E-mail address: perez@ing.puc.cl (J.R. Prez-Correa).
Minerals Engineering 22 (2009) 13501358
Contents lists available at ScienceDirect
Minerals Engineering
j our nal homepage: www. el sevi er. com/ l ocat e/ mi neng
requires a great deal of specic experimental information, not nor-
mally found in an industrial environment. Using simple isotherms
calibrated with plant data, Aminian et al. (1998) obtained good
predictions of the steady states of a test SXEW plant. The mixer
was modelled as an ideal continuous stirred tank reactor, hereinaf-
ter, CSTR, and the settler as two compartments in parallel, an ideal
plug ow and an ideal CSTR. Using a more elaborated steady state
model, Pinto et al. (2004) improved the conguration of an SX Cu
plant applying multi-objective optimization. They included chem-
ical reaction kinetics and a non-ideal hydrodynamics in the mixer
settler units. Then a multi-variable optimizing objective function
was dened that considered equipment geometry, residence time
and agitator speeds, among other design parameters. The optimi-
zation yielded the optimum unit and plant design, which can be
applied for the design of new plants, or for improving existing de-
signs. In spite of being extremely useful in SX plant design, that
model cannot be used to develop and test control systems.
Only dynamic models are useful to design control strategies by
simulation. Most of these models presented in the literature
Nomenclature
A extraction equilibrium isotherm parameter (g L
1
)
a
1A
aqueous active zone fraction of rst CholetteCloutier
unit
a
1O
organic active zone fraction of rst CholetteCloutier
unit
a
2A
aqueous active zone fraction of second CholetteClou-
tier unit
a
2O
organic active zone fraction of second CholetteCloutier
unit
B extraction equilibrium isotherm parameter (g L
1
)
b
1A
aqueous dead zone fraction of rst CholetteCloutier
unit
b
1O
organic dead zone fraction of rst CholetteCloutier unit
b
2A
aqueous dead zone fraction of second CholetteCloutier
unit
b
2O
organic dead zone fraction of second CholetteCloutier
unit
f plug ow fraction of slow owing branch
C stripping equilibrium isotherm parameter
D stripping equilibrium isotherm parameter (g L
1
)
K
E
extraction mass transfer coefcient (h
1
)
K
S
stripping mass transfer coefcient (h
1
)
Q
ai
aqueous mixer inlet owrate (m
3
h
1
)
Q
AL
aqueous owrate in slow ow branch in settler (m
3
h
1
)
Q
am
aqueous mixer outlet owrate (m
3
h
1
)
Q
AUi
aqueous owrate in ith CSTR outlet in fast ow branch
in settler (m
3
h
1
)
Q
oi
organic mixer inlet owrate (m
3
h
1
)
Q
OL
organic owrate in slow ow branch in settler (m
3
h
1
)
Q
om
organic mixer outlet owrate (m
3
h
1
)
Q
OUi
organic owrate in ith CSTR outlet in fast ow branch in
settler (m
3
h
1
)
V
AL1
aqueous volume of rst CholetteCloutier unit (m
3
)
V
AL2
aqueous volume of second CholetteCloutier unit (m
3
)
V
am
aqueous volume in mixer (m
3
)
V
AUi
aqueous volume of ith CSTR (m
3
)
V
OL1
organic volume of rst CholetteCloutier unit (m
3
)
V
OL2
organic volume of second CholetteCloutier unit (m
3
)
V
om
organic volume in mixer (m
3
)
V
OUi
organic volume of ith CSTR (m
3
)
V
S
settler volume (m
3
)
X
*
aqueous equilibrium copper concentration (g L
1
)
X
i
aqueous inlet copper concentration (g L
1
)
X
a
L1
aqueous copper concentration in active zone of rst
CholetteCloutier unit (g L
1
)
X
b
L1
aqueous copper concentration in dead zone of rst Cho-
letteCloutier unit (g L
1
)
X
a
L2
aqueous copper concentration in active zone of second
CholetteCloutier unit (g L
1
)
X
b
L2
aqueous copper concentration in dead zone of second
CholetteCloutier unit (g L
1
)
X
m
aqueous copper concentration in mixer (g L
1
)
X
P
time delayed aqueous copper concentration in slow
ow branch (g L
1
)
X

ST
aqueous equilibrium copper concentration in stripping
stage (g L
1
)
X
Ui
aqueous copper concentration in ith CSTR in fast ow
branch in settler (g L
1
)
Y
*
organic equilibrium copper concentration (g L
1
)
Y
i
organic inlet copper concentration (g L
1
)
Y
a
L1
organic copper concentration in active zone of rst Cho-
letteCloutier unit (g L
1
)
Y
b
L1
organic copper concentration in dead zone of rst Cho-
letteCloutier unit (g L
1
)
Y
a
L2
organic copper concentration in active zone of second
CholetteCloutier unit (g L
1
)
Y
b
L2
organic copper concentration in dead zone of second
CholetteCloutier unit (g L
1
)
Y
m
organic copper concentration in mixer (g L
1
)
Y
P
time delayed organic copper concentration in slow ow
branch (g L
1
)
Y

ST
organic equilibrium copper concentration in extraction
stage (g L
1
)
Y
Ui
organic copper concentration in ith CSTR in fast ow
branch in settler (g L
1
)
e
E
extraction stage organic copper concentration steady
state gradient (g L
1
)
e
S
stripping stage aqueous copper concentration steady
state gradient (g L
1
)
g
E
extraction efciency coefcient
g
S
stripping efciency coefcient
u
A
fraction of aqueous solution owing by slow ow
branch in settler
u
J
fraction of organic solution owing by slow ow branch
in settler
k
1A
aqueous exchange rate between zones in rst Cholette
Cloutier unit
k
1O
organic exchange rate between zones in rst Cholette
Cloutier unit
k
2A
aqueous exchange rate between zones in second Cho-
letteCloutier unit
k
2O
organic exchange rate between zones in second Cho-
letteCloutier unit
Fig. 1. Typical LXSXEW copper plant.
C.M. Moreno et al. / Minerals Engineering 22 (2009) 13501358 1351
include simple thermodynamics and rate expressions, where most
of the complexity is lumped into the hydrodynamics of the settler.
An example of this is the mixersettler cascade model for rare
earths of Wichterlov and Rod (1999). Here, the hydrodynamics
were modelled as a series of CSTRs. However, this model is not that
useful to design control systems since it is too simple, not able to
reproduce the complex dynamics observed in real plants. Wilkin-
son and Ingham (1983) and Ingham et al. (2007) modelled the mix-
er as a CSTR, but included entrainment factors in the outlet
streams. Furthermore, the settler was modelled as two parallel
compartments, just as in the Aminian model. Although this is an
improvement over Wichterlov and Rod model, no comparison
with industrial data has been presented so far. Komulainen et al.
(2006) presented a dynamic model of a SX Cu process calibrated
with plant data, however, the settler is modelled simply as a delay.
Hence, even though the model described the SX process dynamics
well, it will probably be difcult to t that model to other plants
with more complex behaviour.
Here, we present a exible SXCu unit model, able to reproduce
the complex dynamics observed on any stage at any industrial
plant. The model includes McCabeThiele specic diagrams and
complex settler hydrodynamics. The model considers variables
normally measured at industrial facilities and simulates the dy-
namic response of copper concentrations in both phases in the out-
let streams.
2. Process model
Fig. 1 (Jackson, 1986) describes a standard LXSXEW industrial
plant. Leaching (LX) is the rst stage, where copper is extracted by
washing the ore with an acid aqueous solution (SX rafnate). The
outlet stream of LX, the pregnant liquid solution (PLS), is the inlet
stream to the rst stage of the SX process (extraction). In this stage,
the PLS stream is mixed with the barren organic solution (BO), i.e.,
the outlet stream of the second SX stage (stripping), to selectively
transfer copper from the aqueous phase to the organic phase. The
outlet loaded organic solution (LO) from the extraction stage, is
mixed in the stripping stage with the lean electrolyte (LE) stream
coming from the electrowinning process. Since there is a different
pH in the stripping stage, copper transfers now from the organic
phase to the aqueous phase. Then, the outlet aqueous stream from
the stripping stage, rich electrolyte (RE), goes to the electrowinning
stage, where copper is extracted in cathodes, bringing the lean
electrolyte back to the loop.
In this work we developed a general dynamic model for mixer
settlers and t it to two different units, one in the multi-units
extraction stage and the other in the multi-units stripping stage
of an SX process of a LXSXEW copper plant from Molymet S.A.,
similar to that shown in Fig. 1. Although, this plant, with an annual
production of 1 million lb of copper, is rather small compared to
other industrial mining facilities.
2.1. Mass balances
The model was obtained by applying independent copper and
total mass balances on the mixer and the settler (Fig. 2) under
the following assumptions:
Mass transfer in the settler is signicantly lower than in the
mixer, therefore mass transfer is considered only in the mixer
(Wilkinson and Ingham, 1983).
Equilibrium isotherms dene composition gradients for mass
transfer (Steiner and Hartland, 1983).
Perfect mixer (Wilkinson and Ingham, 1983).
Entrainments only occur in the settler (Aminian et al., 1998).
Settler modelled with non-ideal hydrodynamics (Ingham et al.,
2007).
Constant density.
2.1.1. Mixer balances
Applying total mass balances (see Fig. 2) to both, aqueous and
organic phases in the mixer tank, yield,
dV
Am
dt
Q
Ai

Q
Am
Q
Om

V
Am
V
Om
V
Am
1
dV
Om
dt
Q
Oi

Q
Am
Q
Om

V
Am
V
Om
V
Om
2
copper balances differ if the tank is in an extraction or in a stripping
stage. For an extraction unit,
dX
m
dt

Q
Ai
V
Am
X
i
X
m
K
E
Y

E
Y
m
3
dY
m
dt

Q
Oi
V
Om
Y
i
Y
m
K
E
Y

E
Y
m
4
and for a stripping unit,
dX
m
dt

Q
Ai
V
Am
X
i
X
m
K
S
X

S
X
m
5
dY
m
dt

Q
Oi
V
Om
Y
i
Y
m
K
S
X

S
X
m
: 6
In these equations, Q
Ai
and Q
Oi
are the aqueous (subindex A) and or-
ganic (subindex O) mixer inlet owrates, V
Am
and V
Om
are the aque-
ous and organic volumes in the mixer, X
m
and Y
m
are the copper
concentrations in the aqueous and organic phases in the mixer,
respectively, K
E
and K
S
are the copper mass transfer coefcients in
the extraction and stripping units (subindex E stands for extraction
and S for stripping).
X

S
and Y

E
are pseudo-equilibrium copper concentrations de-
ned by empirical equilibrium isotherms (X
*
or Y
*
) and extraction
efciencies (g
E
or g
S
) (Wilkinson and Ingham, 1983).
Y

E
1 g
E
Y
i
g
E
Y

7
X

S
1 g
S
X
i
g
S
X

8
2.1.2. Settler balances
A non-ideal ow (see Fig. 3) for both phases was used to model
the hydrodynamics in this tank (Ingham et al., 2007). The basic
model assumes that the solution is split into two streams, one that
moves fast (fraction 1 u) and the other that moves slowly (frac-
tion u). The fast moving stream, represented by the upper branch
(subindex U) in Fig. 3, is modelled by N perfectly mixed tanks con-
nected in series. Then, in the upper branch, the aqueous phase cop-
per balances are, Fig. 2. Flow diagram of a mixersettler unit.
1352 C.M. Moreno et al. / Minerals Engineering 22 (2009) 13501358
i 1
dX
Ui
dt

1 u
A
Q
Am
V
AUi
X
m
X
Ui
9
i 2. . . N
dX
Ui
dt

Q
AUi1
V
AUi
X
Ui1
X
Ui
10
and the organic phase copper balances are,
i 1
dY
Ui
dt

1 u
O
Q
Om
V
OUi
Y
m
Y
Ui
11
i 2. . . N
dY
Ui
dt

Q
OUi1
V
OUi
Y
Ui1
Y
Ui
12
In turn, the slowly moving stream, represented by the lower
branch in Fig. 3 (subindex L), is modelled as a plug ow tank in ser-
ies with two CholetteCloutier units (Aminian et al., 1998). Each of
these consists of two perfectly mixed zones, an active (top) and a
dead one (bottom). Total and copper balances for both phases in
the plug ow tank are,
Q
AL
t u
A
Q
Am
Ht s
AL
13
X
P
t X
m
Ht s
AL
14
Q
OL
t u
O
Q
Om
Ht s
OL
15
Y
P
t Y
m
Ht s
OL
16
Here, H is the Heaviside step function, which represents a time
delay. Copper balances for both phases in the two CholetteClou-
tier units are given by,
dX
a
L1
dt

Q
AL
a
1A
V
AL1
X
P
k
1A
X
b
L1
1 k
1A
X
a
L1

17
dX
b
L1
dt

k
1A
Q
AL
b
1A
V
AL1
X
a
L1
X
b
L1

18
dX
a
L2
dt

Q
AL1
a
2A
V
AL2
X
a
L1
k
2A
X
b
L2
1 k
2A
X
a
L2

19
dX
b
L2
dt

k
2A
Q
AL1
b
2A
V
AL2
X
a
L2
X
b
L2

20
dY
a
L1
dt

Q
OL
a
1O
V
OL1
Y
P
k
1O
Y
b
L1
1 k
1O
Y
a
L1

21
dY
b
L1
dt

k
1O
Q
OL
b
1O
V
OL1
Y
a
L1
Y
b
L1

22
dY
a
L2
dt

Q
OL1
a
2O
V
OL2
Y
a
L1
k
2O
Y
b
L2
1 k
2O
Y
a
L2

23
dY
b
L2
dt

k
2O
Q
OL1
b
2O
V
OL2
Y
a
L2
Y
b
L2

24
Therefore, the nal copper concentration at the exit of the set-
tler is a weighted sum of both outlet branches,
X u
A
X
a
L2
1 u
A
X
UN
25
Y u
O
Y
a
L2
1 u
O
Y
UN
26
In these equations, a stands for active zone, b stands for dead zone
and k stands for exchange rate between zones.
2.2. Constitutive equations
2.2.1. Equilibrium
The equilibrium isotherm in an extraction unit (see Eq. (7)) is
given by (Komulainen et al., 2006),
Y


AX

B
27
A and B are empirical coefcients tted with process measurements
and extractant manufacturer data as follows:
A a ML 28
B
10
pH

b
PLS
ML
c
d Cu
2
h i
PLS
f Cu
2
h i
BO

29
where ML is the maximum load (g L
1
) of Cu in the organic stream
(extractant manufacturer data), and pH and [Cu
+2
] are daily aver-
aged off-line measured values. In these equations, a, b, c, d and f
are tting parameters (see Table 1).
(1-
A1
) Q
Am1
X
m1
Q
Am1
,X
m1
Q
A1
,X
1
V
UA1
V
UAN
a
1A
V
L1A
X
L1
A
Q
AL1
X
P1
Q
AL1
X
L2
A
Q
AL1
x
L1
A

1A
Q
AL1

1A
Q
AL1

A1
Q
Am1
X
m1

AL
b
1A
V
L1A
X
L1
B
a
2A
V
L2A
X
L2
A
b
2A
V
L2A
X
L2
B

1A
Q
AL1
Q
AUN
X
UN
Fig. 3. Hydrodynamic model of aqueous solution in the settler; the hydrodynamic model for the organic solution is the same, but with sub-indexes O instead of A, and
copper concentrations Y instead of X.
Table 1
Isotherm tting parameters.
Parameters Values
a 0.99
b 1.02
c 1.01
d 35.15
f 27.15
g 0.11
h 444.49
m 0.10
n 0.81
p 8.91
C.M. Moreno et al. / Minerals Engineering 22 (2009) 13501358 1353
Similarly, for the stripping isotherm (see Eq. (8)) (Komulainen
et al., 2006),
Y

C X

D 30
C g ML 31
D h
vol:%
m
H
2
SO
4

n
LE
p 32
where vol.% is the % v/v of reactant in the organic and [H
2
SO
4
]
LE
is
the free acidity of the lean electrolyte (g L
1
); g, h, m, n and p are t-
ting parameters (see Table 1). Variables involved in Eqs. (28)(32)
were measured from common shift samples in the laboratory by
chromatography (see Table 2). Fig. 4a shows daily averages of calcu-
lated isotherm parameters A, B, C and D. Further details regarding
tting isotherm parameters are given in Appendix A.
2.2.2. Mass transfer
In this study, the extraction rate is modelled by a mass transfer
expression obtained from the interface theory (Jackson, 1986),
K
E

Q
Ai
V
Am
X
i
X
m

E
Y
m


Q
Ai
V
Am
X
i
X
m

e
E
33
K
S

Q
Oi
V
Om
Y
i
Y
m

S
X
m


Q
Oi
V
Om
Y
i
Y
m

e
S
34
In these equations, the steady state copper concentration gradients
(e
E
Y

ST
Y
m
for an extraction unit and e
E
X

ST
X
m
for a strip-
ping unit) are tting parameters (Komulainen et al., 2006). Copper
concentrations in Eqs. (33) and (34) (X
i
, Y
i
, X
m
, Y
m
) were measured
in the laboratory by chromatography from shift samples (see Table
2). Fig. 4b shows daily averages of calculated K
E
and K
S
coefcients.
2.2.3. Efciencies
Efciencies for both, extraction and stripping units, are dened
by,
g
E

e
E
Y
m
Y
i

Y
i
35
g
S

e
S
X
m
X
i

X
i
36
Table 2
Measured variables in LXSXEW copper plant.
Measured variables Sampling time (h) Symbols
Inputs
Aqueous inlet owrate 1 Q
ai
[Cu
+2
] aqueous inlet 1 X
i
Organic inlet owrate 1 Q
oi
[Cu
+2
] organic inlet 1 Y
i
Mixer volume 1 V
m
Settler volume 1 V
s
pH of PLS 1 pH
PLS
Free acidity of the lean electrolyte 1 [H
2
SO
4
]
LE
Outputs
Volumes of phases in mixer (i = A, O) 1 V
im
Flowrates at mixer exit (i = A, O) 1 Q
im
[Cu
+2
] in aqueous phase at mixer exit 1 X
m
[Cu
+2
] in organic phase at mixer exit 1 Y
m
-1.2
-0.4
0.4
1.2
0 168 336 504 672 840 1008 1176 1344
S
c
a
l
e
d

v
a
l
u
e
s
t (h)
A B C D
(a)
-1.20
-0.20
0.80
1.80
0 168 336 504 672 840 1008 1176 1344
K

(
s
c
a
l
e
d
)
t (h)
E S
(b)
-0.60
-0.30
0.00
0.30
0 168 336 504 672 840 1008 1176 1344

(
s
c
a
l
e
d
)
t (h) E S
(c)
Fig. 4. Calculated model parameters: (a) Isotherm parameters. (b) Mass transfer coefcients. (c) Efciencies.
1354 C.M. Moreno et al. / Minerals Engineering 22 (2009) 13501358
In the mixers, Y
m
and X
m
are the outlet copper concentrations, Y
i
and X
i
are the inlet concentrations, and Y
*
and X
*
are the equilib-
rium concentrations. Fitting parameters, e
E1
and e
S1
, were set by
trial and error to minimize the sum of squared errors between sim-
ulated and measured outlet copper concentrations. Note that an
efciency 1 means that the respective steady state copper concen-
tration gradient is 0; hence, Y

E
Y

or X

S
X

. Fig. 4c shows daily


averaged calculated efciencies. In these calculations, copper con-
centration values averaged every 8 h in the laboratory by chroma-
tography, were used. Nomenclature provides a complete
description of the symbols used.
Simulations were performed with Simulink

software, using
calculated model parameters (shown in Fig. 4) updated daily, input
variables (shown in Fig. 5) measured hourly, and main model
parameters (see Table 3) which were xed during the whole sim-
ulation run (56 days). Flowrates and pH were measured on-line
with magnetic owmeters (Siemens MAG 3100) and standard
pH-meters, respectively. Copper and acid concentrations were
measured in the laboratory by chromatography. In the actual pro-
cess at Molymet S.A., inlet copper concentration standard devia-
tions ranged between 1 and 2 g L
1
, owrates between 0.03 and
0.3 m
3
h
1
, and pH varied between 0.8 and 1.8. To keep conden-
tiality, no further process details can be disclosed.
3. Results and discussion
First, a parameter sensitivity analysis was carried out to aid
model calibration. After calibration, i.e., setting model parameters
at xed values, simulations were compared with measured outputs
of the actual process, shown in Fig. 6. The same analysis was car-
ried out for both extraction and stripping units.
-4.8
0.0
4.8
9.6
0 168 336 504 672 840 1008 1176 1344
t (h)
PLS LE
(a)
-2.8
-1.4
0.0
1.4
0 168 336 504 672 840 1008 1176 1344
Y
(
s
c
a
l
e
d
)
t (h)
BO LO
(b)
-0.40
-0.24
-0.08
0.08
0 168 336 504 672 840 1008 1176 1344
Q
A
(
s
c
a
l
e
d
)
t (h)
PLS LE
(c)
-0.03
-0.01
0.01
0.03
0 168 336 504 672 840 1008 1176 1344
Q
O
(
s
c
a
l
e
d
)
t (h)
BO LO
(d)
X

(
s
c
a
l
e
d
)
Fig. 5. Scaled measured input variables: (a) Copper concentration in aqueous inlet streams. (b) Copper concentration in organic inlet streams. (c) Aqueous inlet owrate. (d)
Organic inlet owrate.
C.M. Moreno et al. / Minerals Engineering 22 (2009) 13501358 1355
Table 3
Model parameters for extraction unit (E) and stripping unit (S). Sub-indexes A and O stand for aqueous and organic phases, respectively. CC stands for CholetteCloutier units.
Parameters Symbols Nominal values (E) Nominal values (S) Units
Fitting parameters
Fraction of plug ow in lower branch f 0.3 0.4
Steady state concentration gradient e
E
0.7 1.9 g L
1
Fraction going through lower branch (i = A, O) u
i
0.8 0.9
Fraction in CC units active zones (i = 1, 2; j = A, O) a
ij
0.6 0.7
Exchange rate between CC zones (i = A, O; j = 1, 2) k
ij
1.8 1.3
Number of upper branch CSTR tanks N 2.0 2.0
Calculated from tting parameters (see Appendix A)
Upper branch CSTR volume (i = A, O; j = 1, N) V
iUj
5.3 2.7 m
3
Plug ow residence time (i = A, O) s
Li
0.9 1.3 h
Volume of CC units (i = A, O; j = 1, 2) V
iLj
14.2 14.3 m
3
Fraction of dead volume in CC units (i = 1, 2; j = A, O) b
ij
0.4 0.3
-0.90
0.10
1.10
2.10
0 168 336 504 672 840 1008 1176 1344
t (h) Simulated-E Measured-E
(a)
-0.35
0.05
0.45
0.85
0 168 336 504 672 840 1008 1176 1344
t (h)
Simulated Measured
(b)
-0.90
0.10
1.10
2.10
0 168 336 504 672 840 1008 1176 1344
t (h)
Simulated-S Measured-S
(c)
-1.00
-0.60
-0.20
0.20
0 168 336 504 672 840 1008 1176 1344
t (h)
Simulated-S Measured-S
(d)
X

(
s
c
a
l
e
d
)
Y

(
s
c
a
l
e
d
)
X

(
s
c
a
l
e
d
)
Y

(
s
c
a
l
e
d
)
Fig. 6. Simulated vs. measured output variables: (a) Copper concentration in aqueous extraction outlet stream. (b) Copper concentration in organic extraction outlet stream.
(c) Copper concentration in aqueous stripping outlet stream. (d) Copper concentration in organic stripping outlet stream.
1356 C.M. Moreno et al. / Minerals Engineering 22 (2009) 13501358
3.1. Sensitivity analysis
The impact of changing model parameters in the simulation of
output variables was assessed. We were particularly interested in
detecting high impact parameters that could then be used to t
the model to reproduce actual process outputs. Starting from a
nominal case, dened by arbitrarily assigned values of input vari-
ables and model parameters, the model was simulated several
times applying one step change at a time in a given parameter.
By trial and error, we found that for both units (extraction and
stripping), the split fractions, us, and steady state copper concen-
tration gradients, es, were high impact parameters. Therefore, they
were used to calibrate the model for the extraction and the strip-
ping units, running a large number of trial and error simulations.
The rest of model parameters, since they did not affect much sim-
ulation results, were dened after few simulations. The nal model
parameter values are given in Table 3.
3.2. Comparing simulations with measured values
Fig. 6 shows simulated model outputs compared with plant
copper concentration measurements in both phases. It can be seen
that good agreement between simulations and measurements has
been achieved for both units. For the extraction unit, relative mean
squared errors for the outlet stream copper concentrations were
0.03% for the aqueous phase and 6.76% for the organic phase. In
turn, for the stripping unit, relative mean squared errors were
0.07% for the aqueous phase and 2.89% for the organic phase. These
differences in the observed errors are probably due to inherent
sampling and laboratory measurements variability. For example,
measurement errors in the organic phase are larger. In addition,
the extraction unit shows much more variability than the stripping
unit, which is expected since its feed comes directly from the heap
Leaching stage. In turn, this variability is attenuated at the outlet of
the stripping unit due to the damping effect of the mixing tanks.
Despite this difference, the model is exible enough to reproduce
the dynamics of both units well. Furthermore, we compared the
performance of our model against a simple delay model (repre-
sented as a plug ow) for the settler. The simpler model performs
much worse, yielding a relative mean square error of 15.0% for the
aqueous phase (not shown) and 44.8% for the organic phase (see
Fig. 7). Consequently, it is highly benecial to model the settler
with a more complex hydrodynamic model.
Fig. 4 also shows interesting features, like the extreme variation
of the estimated mass transfer coefcients and efciencies. For in-
stance, the signicant drop in the efciency (see Fig. 4c) of the
extraction unit can be attributed to an operational problem and
may be useful to identify its cause. Hence, the model can be rened
more by associating mass transfer and efciencies with specic
operational problems like entrainment, crud formation and vari-
able phase separation times. These would be highly desirable to
improve process supervision by alerting process engineers earlier,
so they can take opportune corrective actions. To develop these
associations though, exhaustive process operation and data analy-
sis are necessary, after running the model in parallel with the pro-
cess for long periods. It is worthwhile noting that the model has a
prediction horizon of 24 h, since calculated parameters (see Fig. 4)
are updated daily based on previous day laboratory data.
The model, as it is can be used to implement on-line state and
parameter estimations, using for example an Extended Kalman Fil-
ter. These estimations can provide further tools to improve process
supervision, and even can help to implement automatic control
strategies. Furthermore, a dynamic model for the entire extraction
SX sub-process can be developed to assess the impact of better
operational practices or alternative automatic control strategies
on process performance, before implementing them in the real
plant, saving time, resources and minimizing the risk of cata-
strophic events.
4. Conclusions
The developed model is a relatively simple form to represent
non-ideal ow in mixer settler tanks. We chose this model struc-
ture since it represents a good compromise between accuracy
and complexity. The main contribution of this work is to show that,
with few sensitive tting parameters, the proposed model provides
realistic dynamic simulations of industrial mixersettler units.
Also, the model can be calibrated with standard measurements
and information normally available in current industrial LXSX
EW plants; therefore, specic dynamic models for a given indus-
trial SX plant can be developed with moderate effort.
Acknowledgements
J.R.P. appreciates the support of AGAUR from the Generalitat de
Catalunya through Grant 2007PIV-00017 and Ponticia Universi-
dad Catlica de Chile for nancial support for a sabbatical stay at
Department DEnginyeria Qumica at Universitat Rovira i Virgili.
C.M.M appreciates the support of Molymet S.A. for providing
experimental information and Ponticia Universidad Catlica de
Chile for granting a PhD scholarship. We appreciate the comments
of the anonymous referees that help us to signicantly improve the
readability of this paper.
Appendix A. Model calibration
A.1. Equilibrium isotherm calculations
For a collection of pH, and [Cu
+2
]
PLS
values, the extractant man-
ufacturer provides several McCabe diagrams, that can be used to
obtain isotherm coefcients, A and B, as a correlation of these vari-
ables. Additionally, plant measurements of ML, pH, [Cu
+2
]
PLS
and
[Cu
+2
]
BO
(normally available at industrial facilities) were averaged
-0.35
0.05
0.45
0.85
0 168 336 504 672 840 1008 1176 1344
t (h)
Model Delay Real
Y

(
s
c
a
l
e
d
)
Fig. 7. Comparison between our model and a simplied pure delay settler model. Copper concentration in extraction organic-outlet stream.
C.M. Moreno et al. / Minerals Engineering 22 (2009) 13501358 1357
daily and, through Eqs. (28) and (29), estimated values of isotherm
coefcients
^
A and
^
B were obtained. Then, parameters a, b, c, d and f
were tted to minimize the following sum of squared differences,
Min
X
56
i1
A
i

^
A
i
h i
2
B
i

^
B
i
h i
2

A1
where subindex i refers to the respective day. The tted values are
given in Table 1.
Similarly, for a collection of [H
2
SO
4
] values, a set of McCabe dia-
grams was given by the manufacturer, yielding a set of values for C
and D. Additionally, a set of plant measurements of [H
2
SO
4
]
LE
and
vol.% were averaged daily and, through Eqs. (31) and (32), esti-
mated values of isotherm coefcients
^
C and
^
D were obtained.
Then, parameters h and p were tted to minimize the following
sum of squared differences,
Min
X
56
i1
C
i

^
C
i
h i
2
D
i

^
D
i
h i
2

A2
A.2. Mass transfer and efciency calculations
Mass transfer and efciency coefcients were obtained using
Eqs. (33)(36). Values of e
E
and e
S
were calibrated by trial and error
after several simulations. The nal values of e
E
and e
S
were those
that minimized the mean squared error between the simulated
and the measured response.
A.3. Calculated parameters in Table 3
They were obtained from,
V
Uij

V
S
1 u
N
A3
s
Li

f uV
S
Q
im
A4
V
Lij

1 f uV
S
2
A5
b
ij
1 a
ij
A6
References
Aminian, H., Bazin, C., Hodouin, D., 1998. Residence time distributions in SXEW
equipment. International Journal of Mineral Processing, 235242.
Bergh, L.G., Yianatos, J.B., 2001. Current status and limitations of copper SX/EW
plants control. Minerals Engineering 14 (9), 975985.
Ingham, J., Dunn, I.J., Heinzle, E., Prenosil, J.E., 2007. Chemical Engineering
Dynamics. An Introduction to Modelling and Computer Simulation. Wiley-
VCH. pp. 174183.
Jackson, E., 1986. Hydrometallurgical Extraction and Reclamation. John Wiley &
Sons. pp. 109138.
Komulainen, T., Pekkala, P., Rantala, A., Jamsa-Jounela, S.L., 2006. Dynamic
modelling of an industrial copper solvent extraction process.
Hydrometallurgy 81 (1), 5261.
Pinto, G.A., Durao, F.O., Fiuza, A.M.A., Guimaraes, M.M.B.L., Madureira, C.M.N., 2004.
Design optimisation study of solvent extraction: chemical reaction, mass
transfer and mixersettler hydrodynamics. Hydrometallurgy 74, 131147.
Steiner, L., Hartland, S., 1983. Unsteady-state extraction. In: Lo, T.C., Baird, M.H.I.,
Hanson, C. (Eds.), Handbook of Solvent Extraction. John Wiley & Sons, pp. 249
264.
Van Bochove, G.H., Krooshoff, G.J.P., de Loos, T.W., 2000. Modelling of liquidliquid
equilibria of mixed solvent electrolyte systems using the extended electrolyte
NRTL. Fluid Phase Equilibria 171, 4558.
Wichterlov, J., Rod, V., 1999. Dynamic behaviour of the mixersettler cascade.
Extractive separation of the rare earths. Chemical Engineering Science, 4041
4051.
Wilkinson, W.L., Ingham, J., 1983. Dynamic behavior and control. In: Lo, T.C., Baird,
M.H.I., Hanson, C. (Eds.), Handbook of Solvent Extraction. John Wiley & Sons, pp.
853886.
1358 C.M. Moreno et al. / Minerals Engineering 22 (2009) 13501358

Você também pode gostar