Вы находитесь на странице: 1из 17

See

discussions, stats, and author profiles for this publication at:


https://www.researchgate.net/publication/238372934

Kinetics of soda ash leaching of


low-grade scheelite concentrates
ARTICLE in HYDROMETALLURGY SEPTEMBER 1996
Impact Factor: 1.93 DOI: 10.1016/0304-386X(95)00099-3

CITATIONS

READS

14

42

1 AUTHOR:
J.I. Martins
University of Porto
49 PUBLICATIONS 734 CITATIONS
SEE PROFILE

Available from: J.I. Martins


Retrieved on: 06 February 2016

hydrometaUurgy
ELSEVIER

Hydrometallurgy42 (1996) 221-236

Kinetics of soda ash leaching of low-grade scheelite


concentrates
J.P. Martins
Departamento de Engenharia Qufmica, Laborat6rio de Cat6lise e Materiais, FEUP, Rua dos Bragas, 4099
Porto, Portugal

Received 7 July 1995; accepted 14 October 1995

Abstract

The reaction kinetics for the leaching of low-grade scheelite concentrates with sodium
carbonate have been studied to establish the effect of the most important process variables, such as
stirring, solid/liquid ratio, sodium carbonate concentration, temperature and particle size. The
leaching residues were chemically and structurally characterized by X-ray diffraction. The
experimental results conform with the shrinking core model and they show that the scheelite
leaching is under diffusion control for about 30 min after the start of the reaction in the range
100-250C. Good tungsten recovery can be obtained.

1. I n t r o d u c t i o n

Dressing techniques, such as gravity concentration, froth flotation, magnetic and


electric fields, are generally effective in the separation of scheelite from associated
gangue, but some of the tungsten concentrates produced still contain only 5 - 2 0 wt%
W O 3. It is thus necessary to treat these concentrates if we want to raise the tungsten
recovery. Taking into account the common mineralogical composition of scheelite ores,
hydrometallurgical alkaline processes are advisable to minimize leaching of impurities.
As is well known [1], the inability to dissolve scheelite in caustic solutions is a result of
thermodynamics rather than kinetics. Otherwise, the Pourbaix diagram for the C a - W CO 3 + H 2 0 system at 25C shows that calcium tungstate is relatively unstable to the
carbonate. Using thermodynamic data [2-4] at 25C, the reaction:
CaWO4(s) + Na2CO3(aq ) ~ Na2WO4(aq ) + CaCO3(s)

(1)

has a standard enthalpy and free energy of 37.6 k J / m o l and 56.5 k J / m o l , respectively,
which means that it is essential to operate at high temperatures.
0304-386X/96/$15.00 Copyright 1996 Elsevier Science B.V. All rights reserved.
SSDI 0304-386X(95)00099-2

222

J.P. Martins / Hydrometallurgy 42 (1996) 221-236

The first report of the use of sodium carbonate to digest scheelite was by Hamilton
[5]. However, only with the work of Masleniskii [6] was the decomposition of scheelite
under pressure discussed. Following this pioneering work, Masleniskii [7,8], Perlov
[8-10] and Zelikman and Radova [11] carried out further investigations. About a decade
ago, Belikov and Masleniskii [12] and Queneau [13] studied the kinetics of dissolution of
scheelite in alkaline solutions. They concluded that the diffusion of calcium or carbonate
ions through the calcite layer formed during the leaching may be the rate-controlling
step. More recently, Zelikman and Meerson [14] have considered that the leaching is
controlled by the rate of chemical reaction. As this process is useful it continues to
attract the attention of researchers; improvements, such as the simultaneous action of
other alkaline compounds with sodium carbonate [15,16], counter-current extraction of
concentrates [17] and other means of increasing process efficiency [18-20] have been
proposed.
The aim of this work is to establish the effect of the process variables on the reaction
rate of soda ash leaching of low-grade scheelite concentrates and the use of a
counter-current extraction process to improve tungsten recovery.

2. Non-catalytic solid-fluid reaction models


There are several kinetics models to describe non-catalytic solid-fluid reactions
[21-27]. The measurement of the specific surface area of the low-grade concentrate after
grinding to < 100 txm indicated a surface area of < 0.05 m2/g. This suggests that the
solid particles are practically non-porous. The chemistry and physical examination of the
pellet after leaching show the presence of calcium carbonate surrounding an unreacted
scheelite core. Therefore, the unreacted scheelite core during leaching moves from the
outside towards the centre; without any cracking or swelling of the product film the size
of the pellet is maintained. The shrinking core model [23] seems to be a suitable
mathematical model to describe the kinetics of the process.
2.1. Conversion expressions f o r the shrinking core model

Considering the approximation to the pseudo-stationary state, the evolution of the


unreacted core with time for first and zero reaction orders, respectively, to the leaching
agent and to the scheelite, is given by the relation:

R s[ (1

t= CA---~

KmA

R),

DeA

1 -- r a / R 3) + --psKs 1 -

+ ~

(1 - r 2 / R 2)

(2)
which for the total conversion, t = -r and r e = 0, becomes:
Rps(
1
1
R )
" r = ~ A 3Km-"-'~+ ~KsPs + ~6DeA

(3)

J.P. Martins/Hydrometallurgy 42 (1996)221-236

223

If the fluid film resistance controls the reaction rate, Kma << K s, DeA , and the ratio
between Eq. (2) and Eq. (3) gives:
t
- = X with "r = psR//3KmACAo
I"

(4)

If the reaction rate is independent of the fluid film resistance, KmA ~ ~, and the ratio
between Eq. (2) and Eq. (3) gives:

. . . .
-r
0c+6

1-(l-X)

'/3] + ~ [ 1

+2(1-X)-3(1-X)

2/3]

(5)

with "r = R(0c + 6)/6KsCAo; where 0 is a non-dimensional variable that expresses the
relative importance of the specific rate and the diffusion through the ash layer product.
When 0~ ~ 0, the chemical relation controls the reaction rate and Eq. (5) becomes:
t
-

'I"

= 1 - (1 - X) '/3 with 'r = Rps/KsCAo

(6)

and when 0~ ~ ~, the ash layer controls the reaction rate and Eq. (5) becomes:
t
- = 1 + 2(1 - X) - 3(1 - X) 2/3 with "r = R2ps/6DeACAo
T

3. E x p e r i m e n t a l

(7)

details

3.1. Characterization of scheelite


The low-grade concentrate was supplied by Minas de Tarouca. Chemical analysis, by
atomic absorption spectroscopy and gravimetric methods, mineralogical composition, by
X-ray powder diffraction pattern, and sieve analysis are shown in Tables 1 and 2.

Table 1
Chemical analysis and mineralogical composition of the low-grade scheelite concentrate
Chemical analysis
Mineralogical composition
Chemical constituent

Content (%)

Mineral

Content (%)

WO3
SiO2
CaO
Fe203
AI203
S
As
Na20 + K20
MgO
MnO
Sn
Loss on pyrolysis

5.80
37.50
16.98
13.08
11.22
4.98
1.87
1.53
0.90
0.34
0.05
5.13

Scheelite
Calcite
Apatite
Quartz
Mica, caulinite and feldspar
Pyrrhotite
Arsenopyrite
Cassiterite

7.80
23.46
1.90

26.27
25.27
11.48
3.96
0.06

J.P. Martins/ Hydrometallurgy42 (1996) 221-236

224

Table 2
Sieve analysis of three lots of low-grade scheelite concentrate
Size, mesh

Fraction (%)

140
200
325
400
< 400
Median size (ixm)

Lot A

Lot B

Lot C

21.3
22.8
23.7
6.4
25.8
60

22.1
21.1
56.8
40

30.3
17.5
52.2
40

3.2. Leaching trials


The leaching trials at atmospheric pressure were carried out in Pyrex Quickfit vessels
thermostatically controlled at 100 + IC and under pressure in a 1 1 Parr model 4531
autoclave, using 2 0 - 8 0 g o f sample. The time o f experiment was taken after attaining
the required temperature. A t the end of a run, the heating was stopped and the autoclave
allowed to cool to 70C.
Extraction was carried out using two types o f systems: a continuous counter-current
system - - I (Fig. 1) - - and a semi-continuous counter-current system - - II (Fig. 2).
F r o m each o f the processes under steady-state conditions we can conclude that the
number of solid and fluid interactions are five for system II and three for system I. The
Na2CO 3 used in the experiments was of reagent grade.
A t the end o f each experiment the mixture was filtered and the residues washed,
using about twice the volume o f the filtrate. Aliquots were collected and analysed for
tungsten by the mercurous method and a volumetric method was used for carbonate.

4. Results and discussion


4.1. Effect of stirring
The stirrer used was the pitched blade turbine; other details are: stirrer system are:
i m p e l l e r / v e s s e l diameter ratio 0.6; number o f stages 2, with a spacing of 0.04 m; and
h e i g h t / d i a m e t e r impeller ratio 0.1. The results, presented in Table 3, indicate that the
stirring rate has little effect on the dissolution rate of scheelite when the speed is
changed from 300 min - l to 800 min -1. It is therefore reasonable to conclude that
beyond 750 m i n - t the solid particles are not kinetically liable to diffusion control by the
adherent fluid film.

So=fresh sol.ution
Sl
$2
S3=#ina(.(sotution
~l lgStage j-J 2eStage j=J 3aStage jtoo=fresh ore
n~= Leached ore
m2
mI
Fig. 1. Scheme of a continuous countercurrent system.

J.P. Martins/Hydrometallurgy 42 (1996) 221-236

225

1a Stage: initially, reactors A and B in


operation and C discharge
l-to set the fresh ore inside C
2-change the leaching solution fi-om B to C
3-change the leaching solution from A to B
4-to bring fresh leaching solution to C

3x

2'~ Stage:
1-separation of solids and liquids at C and
leave of a solution with two stages of
leaching
2-change the leaching solution from B to C
3-change the leaching solution from A to B
4-removal of sterile leached ore from A

t~

3~dStage:
1-to set the fresh ore inside A
2-change the leaching solution from C to A
3-change the leaching solution from B to C
4-to bring fresh leaching solution to B

=Lk,.~j

4th Stage:
l-separation of solids and liquids at A and
leave of a solution with three stages
of leaching
2-change the leaching solution from C to A
3-change the leaching solution fi'om B to C
4-removal of leached ore from B
5Ih Stage:
1-to set the flesh ore inside B
2-change the leaching solution from A to B
3-change the leaching solution from C to A
4-to bring fresh leaching solution to C

-1~ 2

6~ Stage:
1-separation of solids and liquids at
leave of a solution with five
of leaching
2-change the leaching solution from
3-change the leaching solution from
4-removal of leached ore from C

B and
stages
A to B
C to A

Fig. 2. Scheme of a semi-continuous countercurrent system.

4.2. Effect of solid/liquid ratio


The influence o f s o l i d / l i q u i d ratio was studied at 100C for a leaching time of 3 h
using different concentrations o f Na2CO 3. Fig. 3 shows that, for the same stoichiometric

J.P. Martins / Hydrometallurgy 42 (1996) 221-236

226

Table 3
Influence of stirring rate on the conversion of the scheelite
Time

Na2CO 3 / W O 3

Stirring rate (rain- 1)

(h)

(mol/mol)

300

500

750

800

1.00
1.00
1.00
2.25
2.25
2.25
4.50
4.50
4.50

2
4
8
2
4
8
2
4
8

14.01
35.22
41.19
18.36
43.71
52.98
18.84
50.26
67.24

14.66
34.18
42.80
19.05
44.22
53.75
20.95
53.10
67.24

18.42
35.18
43.32
23.04
43.81
52.52
23.43
53.40
67.85

18.30
35.20
43.12
23.15
44.20
53.05
23.12
53.30
67.60

Experimental conditions: p = 1 atm; T = 100+ IC; solid/liquid ratio = 1/5; ore = lot A ( < 100 Ixm).

relation of Na2CO3/WO 3 (solid lines), the conversion increases when the solid/liquid
ratio increases, which shows a kinetic dependence on the concentration of the leaching
agent. It is also clear for the same concentration of sodium carbonate (dashed lines), that
there is an increase in the conversion with a decrease in the solid/liquid ratio. The
maintenance of sodium carbonate concentration with a decrease in the solid/liquid ratio
demands a larger excess of the reagent over the stoichiometric requirements. Therefore,
the bulk concentration can be considered unchangeable with time, which confirms the
hypothesis of the pseudo-stationary state.

4.3. Effect of sodium carbonate and temperature


The plotting of the correlation conversions against time, X = fit), shows two distinct
zones (Figs. 4-7): in the first few minutes, the slope of the function is constant and
increases with the temperature; for longer leaching times the reaction rate decreases and

50

hlfl

~o
.[

30

u 1l
1/3

1/5
117
solid/liquid ratio

111(

Fig. 3. Correlation between the yield of WO 3 leaching and solid/liquid ratio for different concentrations of
Na2CO 3. Experimental conditions: time = 3 h; stirring rate = 750 m i n - t ; ore = lot A( < 100 mesh); temperature = 100C.

J.P.

Martins/Hydrometallurgy42 (1996) 221-236

227

100
~9G

e=~1

..3

~/*0 /# ~
20

l ternperature.100ZlO C

.I,.I

l,

sti rring-?50min-I

ore=tot A

I I8I 19

5 6
time(h)

10 11

Fig. 4. Effect of sodium carbonate concentration on the percentage of W O 3 extracted at 100C for lot A.

attains a constant value. It is possible to say that a maximum recovery of about 80 wt%
WO 3 is gained about 2 h after the start of the reaction at 100C, after 1 h at 150C, after
30 min at 200C and after 15 min at 250C. The calculation of apparent activation
energy with the Arrhenius equation provides a further criterion for determining the
controlling step of the reaction. The Arrhenius plot constructed for the first 15 min and
30 min of the reaction, Fig. 8, gives, respectively, an apparent energy activation of 41
kJ/mol and 18 kJ/mol. These values indicate kinetic control at the start of the reaction,
which changes to diffusion control. Therefore, the ash layer of the calcium carbonate,
associated with the structure of the solid particles, has a particular role in the leaching
process. The X-ray powder diffraction pattern on the external surface of a grain after
leaching is reproduced in Fig. 9. The calcite detected confirms the presence of an

100

_,0(
o/;

r..-1

6O

f
I

o
u

20

I
temperature

= lS0-~50C

sot i d/liquid ratio=lllO


stirring = 750 min-t

3
time ( h I

i
!

Fig. 5. Effect of sodium carbonate concentration on the percentage of WO 3 extracted at 150C for lot A.

J.P. Martins / Hydrometallurgy 42 (1996) 221-236

228
100

f
8O

o+
o

11.5 I

p
____----4-

60

P
I

!
l
1!=2

f ~ ' ~
/"

20

t 11mpllratur11= 200.+5+C
sotid/|iquid rati0 = 1110
stirring= 750min-I

i r

J
0

3
t i m e [h)

&

Fig. 6. Effect of sodium carbonate concentrationon the percentage of WO3 extracted at 200"C for lot A.

insoluble film on the unreacted core during leaching, as has also been described in the
literature [8,28].
The analysis of the impurities in some leach liquors, Table 4, shows the purifying
effect o f alkaline leaching solutions on low-grade scheelite concentrates. Yield can be
defined as the moles of a particular product per mole of initial leaching reactant.
Therefore, the ratio of the useful product to the yield of secondary products is the
selectivity, which is the ratio o f co-ordinate y axis to the co-ordinate x axis in Fig. 10.

100

e= 4

.~80
C
~60

20

e=3

r---"-'-

t 11mp'11rature: 250~5%
$olid/tiquid ratio= 1/10
stirrlng=?50 rain-+

:~
time(h)

~,

Fig. 7. Effect of sodium carbonate concentration on the percentage of WO 3 extracled at 250C for lot A.

J.P. Martins/Hydrometallurgy 42 (1996) 221-236

229

.=,
<~-

) ~

'q

c5

i od,:Sd~

t",i

~"~

e,'~

0
0
V

o
"~'=

<

< .=.
0
...:"

',O

e,i

~ ,-.: ,.~,::5 d

e,i
II

e.
o
o

~"

I 0

.,I

.~

r:

6
0

o.o

o
e

z
e.

.,,..

e~
o

.o
"a
0
e.
-

o
"+t
0

e.

t'-,i

~ o

e4,.-: d

..-:

t..-.
~A
o

:=,
o

<

,-.

J.P. Martins / Hydrometallurgy 42 (1996) 221-236

230

23

21

f
19

' 1'7

15

13
1.8

_.c

No2 COz- i0.619/t


sotid/t~quid r a t i o - I / 1 0
stirring- 750 min-I
1 - - t - 1 5 m i n ~ E - 41 KJ / m o t
2 - t - 3 0 m i n c E - 18KJ /rnoL

2,0

2.2

2/,

2.6

2.8

103~(1/T)1 K-~ )

Fig. 8. Arrhenius plots for the dissolution of scheelite for the ore of lot A ( < 100 mesh). D = l (t = 15 rain);
O = 2 (t = 30 min).

From it we see the decrease in sodium carbonate selectivity to the leaching of scheelite
with increasing concentration. For temperatures higher than 100C, the concentration of
silicon varies between 0.5 and 1.3 g / l , of which 0.3 g/1 remain in solution after
cooling. Increasing the temperature from 100C to 250C results in sodium carbonate
loss in secondary reactions increasing from 8 wt% to 24 wt%.

4.4. Effect of particle size


Ore was crushed and classified by a standard sieve in three lots: A ( < 150 Ixm), B
( < 75 ixm) and C ( < 53 Ixm). Their distribution size fractions, Table 1, show for lot B
and lot C the same median size, so only lot B was used to compare with lot A results.
Fig. 11 illustrates the leaching results at 100C for lot B. The comparison of these results

t~
U

3~.,.

'

2~.4

<

2~.4

20

Fig. 9. X-my power diffraction pattern (single-crystal monoc~omatized CuK eL radiation).

J.P. Martins / Hydrometallurgy 42 (1996) 221-236


2.0

231

1.6

0.&

temperature = 100t1*C
s o t i d / | i q u i d ratio=l/5
s t i r r i n @, 750 min-t
ore,tot'B

0.4

[--

0.6

1.2

I CA~ CA-Cc )*10


CAo
Fig. lO. The analysis of selectivity of sodium carbonate at 100C.

with those of lot A, Fig. 3, shows an increase of about 4 wt% in the recovery of tungsten
with a sodium carbonate concentration of 15.9 g/1 and 10 wt% in the recovery of
tungsten for higher concentrations of leaching reagent.

4.5. Effect of counter-current

extraction

The results obtained with continuous and semi-continuous systems are shown in Fig.
12 and we can compare the values obtained in a single stage of leaching for the same
stoichiometric Na2CO3/WO3 ratio and residence time. The improvement in tungsten
recovery is associated with a reduction in WO 3 lost in the gangue. This behaviour

100
9O

~7c
eSC

rf . ~ e=2

~3C
~2C

stir ring ='/SOmin-1


,-1

10
1

t e mperat ure=100:l*C

solid/tiquid ratio,l~

~-4 "'"

ore=totB

IIII

& 5 6 7
time(h)

9 10 11

Fig. 11. Effect of sodium carbonate concentration on the percentage of WO 3 extracted at 100C for lot B.

J.P. Martins/Hydrometallurgy 42 (1996) 221-236

232

L,

time(h)
12

120

.~,

c~100

~ .-J = , . - - . - - - . . - - ~ - -

~ 80
,e.-S... _ .~ . . . .

60

.o~.~
,one stage of leaching
~"ko~''l~
muttiple stages of leaching
t~t I~.rat ur~-1q0~1~..
a)semicontinuous system
i(Vtiquie r~tm.l/',
blcontinuous system
st irring ='FJOmin-1
I
I
t
I
I
1
2
3
/,
5
s rages of teaching
~,.,.~-

/,0

Fig. 12. Results of counter-current extraction for continuous and semicontinuous systems at 100C.

confirms the principles of reversibility of Eq. (1). In fact, the contacts of a solid
successively more exhausted and a solution less concentrated in sodium tungstate aid the
displacement of Eq. (1) to the right.
4.6. Shrinking core model

The values of the apparent energy activation suggest diffusion control through the ash
layer in accordance with the results of applying Eq. (7) of the shrinking core model (Fig.

so.tid/liquL& rat i o.1/10


S t I rrJnaul~Umlfl'1
ore= to~,~
I

.=

e,

/
J

=4

0
I~

Oc G

[(1-(1-X)1/3 ] . _~_[12(1_X)_3(1.X)2/3]]=lO

Fig. 13. Appfication of the shrinking core model to the leaching of scheelite with Na2CO 3 (7.96 g/l),
considering diffusion control in grains.

J.P. Martins / HydrometaUurgy 42 (1996) 221-236

233

10

//

e,3

.c

~4

/,!

//
rj ~ernpe rat ure-lO0~l*C

/ temperature=lOOtl*C
sotid/liquid ratio=l/5
,P st irring.750 min-t

~,"

f solidAiquid ratio-l/5

stirring.750 rnin -1
ore-tot B

...~

ore-to'tA I

2
Z.10

Z.IO

[ 1 . 2 (1 -X ) -3 (1-X)2/3] = Z

Fig. 14. Application of shrinking core model to the leaching of lots A and B of scheelite for two
concentrations of sodium carbonate, considering difusion control in grains.

13). For an Na2CO 3 concentration of 7.96 g/1 the experimental results also show that
increasing the temperature brings forward the beginning of that control: 60 min at
100C; 30 min at 150C; 20 min at 200C; and 10 min at 250C.
According to the proposed model, the influence of particle size is given by:
T1 ~--

R2t / R 22

(8)

T2

for total conversion of particle size Rl; 'I"2 = time for total conversion
of particle size R E.
Calculating x from the experimental results for the leaching of lots A and B, the
slope of the functions shown in Fig. 14, and substituting in Eq. (8), ratios of 2.1 for
[Na2CO 3] = 15.91 g / l and 2.0 for [Na2CO 3] -- 21.21 g/1 are found. These values are
acceptable since the theoretical value is 2.25, based on the median size of the two lots of
the solid particles.
Considering the influence of temperature on the diffusivity of the fluid reactant using
the relation:
where: T1 = time

DeA = DoTm

(10)

we obtain for "r the following expressions:


"r = c - T - m

R2ps
with c = - 6DoCAo

(11)

and:

log "r = l o g c - m l o g T

(12)

234

J.P. Martins / Hydrometallurgy 42 (1996) 221-236


i

1.3!
!

1.2

t~
=
1.1

No 2 CO~= "7.96 g / L
sotid/Liquid rotio=l/10
stirring-750 minq
ore - [o~ A

~,0
2.5

2.6

tog T

2.7

( K "1 )

Fig. 15. Correlation between "r and temperature.

Taking the slope of the functions from Fig. 13 and correlating them with the
temperature, Fig. 15, we have calculated the unknown quantities D O and m; substituting
into Eq. (10) we get:

DcA=

4 10 - ] 6 . T 0"7 ( m 2 / s )

(13)

which suggests that the effective diffusivity, DeA, may be considered to lie between the
molecular diffusion regime and the Knudsen diffusion regime. This behaviour agrees
well with the porous structure of the ash layer surrounding the unreacted core.
The rate of decomposition of solid particles expressed in terms of the rate of reaction
of the fluid is given by:

v = Ps(-- dt ] = bKsCA

(14)

where: Ps is the molar density of the solid; b is the stoichiometric coefficient; K s is the
heterogeneous rate constant; CAo is the fluid reactant bulk concentration; and rc is the
distance co-ordinate perpendicular to the solid surface.
The specific reaction rate for different temperatures and the apparent activation
energy were obtained with the data of the function X = f(t) for the first 15 min;
expressing the data of the straight line 1 in Fig. 8 in the form of the well-known
Arrhenius equation we get:
K s = 2.7 10 -3" e -5/T ( m / s )

(15)

The overall extent of the reaction is given by:

X= 1 -

(16)

and differentiating with respect to time gives:

( dr]l- - ~ " ( 1 - X )
-- dt

(dX)

-2/3. ~

(17)

J.P. Martins / Hydrometallurgy 42 (1996) 221-236

235

Under diffusion control we can take (dX/dt) from Eq. (7) and substituting in Eq.
(17) we get:
(

drc]
dt ]

DeACAo

=--xRps

[(I - X ) ' / 3 -

(I - X ) 2/3]

(18)

from which with the Eq. (13) allows us to express the reaction rate for lot A by:

v=0st-

( dro
)

= 1.33 1 0 - " "T'TCAo/[(1- X ) ' / 3 - ( 1 -

X)2/3] (kmol/m2. s)

(19)

5. Conclusions

The leaching of low-grade scheelite concentrates by sodium carbonate is strongly


dependent on temperature. However, after a short reaction time the ash layer of calcium
carbonate associated with the structure of the solid particles plays a principal role in
leaching. The experimental results conform with the shrinking core model and the
constant rate and effective diffusivity are found to be: K s = 2.7 10 -3 e -5/T (m/s)
and IDeA -----4 X 10 -16" T 0'7 (m2/s).
The selectivity of the reaction decreases with the temperature and sodium carbonate
concentration, and the loss of tungsten in the ganguc can be reduced using counter-current extraction.

6. List of symbols

IDeA
CAo

CA
Cc
Ks
R
t
Ps
K~
r~
,'r

X
O~
e

effective diffusivity of fluid (m2/s)


initial concentration of fluid reactant A (kmol/m 3)
concentration of fluid reactant A (kmol/m 3)
concentration of product generated C (kmol/m 3)
rate constant (m/s)
radius of spherical initial particle (m)
time (s)
molar density of solid reactant (kmol/m 3)
mass transfer coefficient by convection (m/s)
radius of the unreacted core (m)
time for the total conversion (s)
conversion, 1-(re/R) 3
non-dimensional variable, RKsps/D,A
equivalent number, stoichiometric relation Na2CO3/WO 3 (kmol/kmol)
stoichiometric coefficient in Eq. (1), CaWO4/Na2CO 3 (kmol/kmol)

236

J.P. Martins / Hydrometallurgy 42 (1996) 221-236

References
[1] Osseo-Asare, K., Solution chemistry of tungsten systems. Metall. Trans. B, 13 (1982): 555-564.
[2] Dean, A.J., Lange's Handbook of Chemistry. Mc Graw-HiU, 12th ed. (1979).
[3] Barany, R., Heats and free energies of formation of calcium tungstate, calcium molybdate and magnesium
mobybdate. U.S. Bur. Mines, Rep. Invest., 6143 (1942).
[4] Latimer, W.M., Oxidation Potentials. Prentice-Hall, Englewood Cliffs, N. J., 2rid ed. (1959).
[5] Hamilton, E.M., U.S. Pat. 1,261,383 (1918).
[6] Maslenitskii, I.N., Autoclave process of tungsten extraction from its concentrates. Tsvet. Metal., 4-5
(1939): 140-143.
[7] Maslenitskii, N.N., Soda process for treatment scheelite concentrates. Obogashch. Rud, 2(4) (1957):
3-10.
[8] Maslenitskii, N.N. and Perlov, P.M., Development of the autoclave soda process for the treatment of
tungsten concentrates. Proc. Int. Mineral Processing Congr. Group III, Pap. 41, London (1960).
[9] Perlov, P.M., Interaction of calcium tungstate with soda carbonate in the autoclave process. Obogashch.
Rud, 3(1) (1958): 25-34.
[10] Perlov, P.M., The hydrolyse of Na2CO 3 during treatment of autoclave scheelite. Obogashch. Rud, 3(2)
(1958): 32-34.
[11] Zelikman, A.N. and Rakova, N.N., Condition of decomposition scheelite and powellite with ammonia
solution. Tsvet. Metal., 39(3) (1966): 57-60.
[12] Belikov, V.V. and Maslenitskii, I.N., Kinetics of soda leaching scheelite. Obogashch. ROd, 10(4) (1965):
20-24.
[13] Queneau, P.B., The kinetics of the dissolution of scheelite in basic aqueous solutions. Ph.D. Thesis, Univ.
Minnesota (1967).
[14] Zelikman, A.N. and Meerson, G.A., Metalhirgiya Redldkh Metallov. Metallurgiya, Moscow (1973).
[15] Perlov, P.M., Combined action of solutions of certain compounds on scheelite in the autoclave process.
Obogashch. Rud, 3(21) (1976): 27-29.
[16] Zelikman, A.N., New process in hydrometallurgy of tungsten and molybdenum. Zh. Vses. Khim. Ova.,
16(5) (1971): 548-555.
[17] Martins, J.P., The hydrometallnrgy in the valorization of tungsten ores. Ph.D. Thesis, FEUP/Univ. Porto
(1983).
[18] Quenean, P.B., Huggins, D.K. and Beckstead, L.W., Soda ash digestion of scheelite concentrates. Proc.
Symp. Extraction Metallurgy and Refractory Metals (1981), pp. 237-267.
[19] Rumyantsev, V.K., Makarov, A.N., Polyakov, B.I. and Parfiev, N.A., Continuous separation of a tungsten
concentrate scheelite in an autoclave under pressure. Khim. Redk. Tsvetn. Metal., (1975): 66-68.
[20] Queneau, P.B., Beckstead, L.W. and Huggins, D.K., U.S. Pat. 4,325,919 (1982).
[21] Levenspiel, O., Chemical Reaction Engineering. Wiley, New York, 2nd ed., Chap. 12 (1972).
[22] Habashi, F., Principles of Extractive Metallurgy. Gordon and Breach, New York, Vol. 1, Chap. 7 (1980).
[23] Yagi, S. and Kunii, Fhiidized-solid reactors with continuous solid feed (II). Chem. Eng. Sci., 16 (1955):
372-379.
[24] Wen, C.Y., Non-catalytic heterogeneous solid fluid reaction models. Chem. Eng. Sci., 60(9) (1968):
34-54.
[25] Park, J.Y. and Levenspiel, L., The crackling core model for the reaction of solid particles. Chem. Eng.
Sci., 30 (1975): 1207-1214.
[26] Szekely, J. and Evans, J.W., A structural model for gas-solid reactions with a moving boundary. Chem.
Eng. Sci., 25 (1970): 1091-1107.
[27] Martins, J.P., Cracking shrinking model. Int. Rep. Fac. Eng. Univ. Porto Dep. Eng. Quim., Univ. Porto
(1991).
[28] Fieberg, M.M. and Coetzee, C.F.B., The recovery ot tungsten from South African low-grade scheelite
concentrates. MINTEK Rep. MI64D (1984).

Вам также может понравиться