CoUoids and Surfaces, 15 (1985) 335-353
Elsevier Science Publishers B. V., Amsterdam - Printed in The Netherlands
EFFECfS OF DISSOLVED MINERAL SPECIES ON THE ELECfRO-
KINETIC BERA VIOR OF CALCITE AND APATITE
J. OFORI AMANKONAH and P. SOMASUNDARAN
School of Engineering and Applied Science, Columbia University, New York, NY 10027
(U.S.A.)
(Received 7. May 1984; accepted in final form 22 March 1985)
ABSTRACT
The interfacial behavior of salt-type minerals such as calcite and apatite is governed
to a large extent by their electrochemical properties which in turn are governed by pH
and concentration of dissolved mineral species. In mixed mineral systems, the interfacial
behavior of various minerals is often quite different from what might be expected on
the basis of their individual properties. Predictions for flotation separation based on
single mineral tests often fail in these systems mainly owing to the interactions between
dissolved mineral species.
The effect of dissolved mineral species of the calcite-apatite system is investigated
in this study using the electrokinetic technique. The electrochemical, dissolution and
spectroscopic properties of each mineral were markedly altered by the supernatants of
the other mineral. Most interestingly, under the test conditions, the isoelectric point
of each mineral in the supernatant of the other was observed to be similar to the super-
natant mineral source. It is clear that, under these conditions, a flotation separation
scheme designed on the basis of the surface properties of the single minerals is not likely
to perform satisfactorily. Tests in inorganic solutions have shown that surface reaction
and/or bulk precipitation could be responsible for the observed shifts. Surface chemical
alterations predicted from theoretical considerations using species distribution diagrams
are correlated with the experimental results.
INTRODUCfION
Separation of phosphates from carboni tic gangues by flotation has been
a major problem in the phosphatic industry for years. Direct application
of various processes described in the literature to phosphate ores has often
failed to reduce the carbonate content of the final product to an acceptable
level [1, 2]. Furthermore, a previous investigation [3] has shown predictions
of selective flotation on the basis of single mineral tests to fail even for
synthetic mineral mixtures.
The difficulties encountered in the separation of calcite-dolomite type
impurities from phosphates have been attributed to the similarities in the
surface chemical, electrokinetic and dissolution properties of these minerals
0166-6622/85/$03.30
@ 1985 Elsevier Science Publishers B. V.
336
as a result of which both the carbonates and phosphates respond similarly
to anionic and cationic collectors [4, 5].
The dissolution characteristics of the minerals can be expected to play
a major role in determining the nature of the interactions taking place in
the bulk solution or in the interfacial region and hence the efficiency of
such interfacial processes as flotation and flocculation. In salt-type mineral
systems where the solubility is markedly higher than in most other systems
such as oxides and silicates [6], this role is expected to be even more sig-
nificant.
Effects of dissolved species in various mineral systems have been reported
by several investigators. Healy [7] observed drastic changes in the point of
zero c~e of magnetite when stored in glass containers. The observed
changes were attributed to the silicate species released from the glass con-
tainer. In a similar manner, the dissolved species can cause significant activa-
tion or depression of flotation under certain solution conditions. These ef-
fecw have been correlated with both the specific adsorption of hydroxy
complexes of the cations as well as the stronger tendency of multivalent
ions over monovalents to compete with the surfactant ions for the adsorp-
tion sites [8-14]. Fuerstenau and Miller [15] have obtained good correla-
tion between the pH range in which quartz flotation, using anionic sur-
factants, is activated by cations and the pH range in which the first hydroxy
complex of the cation is formed. This illustrates the role of the specific
adsorption of these complexes. The high adsorption affinity of partially
hydrolyzed ions on charged surfaces has been recognized by Wolstenhamme
and Schulman [16]. Dissolved ferric and other metallic species have been
reported to activate the flotation of silica using fatty acid as the collector
[11, 12, 14, 15]. This suggests that the dissolved species from hematite
can affect silica flotation in hematite-silica systems. It should prove useful
to study the effect of hematite supernatant itself on silica flotation. In a
separate study by Iwasaki and Soto [17], the effect of calcium and mag-
nesium species released into the solution by the impurities associated with
hematite have been shown to have a significant effect on the selective
flocculation of hematite from quartz. This study clearly shows that the
water in equilibrium with the ore can contain sufficient amoun~ of calcium
and magnesium ions to significantly alter the surface properties of silica.
As has been shown by several investigators [18-21], surface precipitation
of a different form of the mineral than its original one from a solution that
is supersaturated is also possible. This fact has been considered in detail by
Parks [22] for alumina systems and by Brown and co-workers [19,20,23]
for phosphates.
Due to the complex nature of salt-type mineral systems, much controver-
sy exists in the literature on the role of polyvalent ions. Polkin [24] con- ,
siders the polyvalent ions to be essential for the formation on hydrophobic
multilayers on calcite. Sun et al [25] and Eigeles [26] consider such ions to
be harmful due to precipitation as metal soaps in the pulp.
337
Surface or bulk precipitation of a more stable phase is perhaps one of
the major effects due to dissolved species. For example, in the presence
of 10-4 kmol m-3 calcium, MacKenzie and Mishra [27] observed that the
apatite surface closely resembled that of the calcite. This could be due to
the precipitation of calcium carbonate on the surface of the apatite. The
implication of this observation is that a separation scheme based on the
interfacial properties of the individual minerals is not likely to perform
satisfactorily for beneficiation of ores. The observation by Le Bell and
Lindstrom [28], and Calara and Miller [29] on the electrokinetic behavior
of fluorite after various treatments must be noted. These authors observed
drastic changes in the point of zero charge of fluorite when it was con-
ditioned in carbonate solutions.
The effects of dissolved mineral species of calcite and apatite on the
flotation of these minerals has been shown recently in our laboratory to
be drastic under certain solution conditions [6]. Flotation of calcite and
apatite in mineral supernatants was drastically depressed due to the pre-
sence of dissolved species.
The electrokinetic behavior of calcite and apatite in water and in mineral
supernatants is discussed in this work. Experiments have also been con-
ducted in various electrolytes containing mineral constituent ions in order
to identify their role when present in mineral supernatants. The observed
changes in the electrokinetic behavior have been correlated with thermo-
dynamic predictions.
EXPERIMENTAL MATERIALS AND METHODS
Materials
Calcite
The calcite of lO-micrometer size was prepared during an earlier investi-
gation by Somasundaran and Wang [4, 30] and was used in this study.
Apatite
Synthetic hydroxyapatite was used in this study. It was prepared by
the precipitation technique described by Moreno et al. [20] and modified
by Somasundaran and Wang [4, 30]. The method involves mixing appro-
priate amounts of K2HPO4 in a KOH solution and Ca(NO3h in water and
boiling the reaction products under appropriate solution conditions to
avoid the formation of octa-calcium phosphate as a precursor. To avoid
the possible precipitation of CaC~ in the system or the formation of car-
bonate species in the crystal lattice of the apatite, the experimental set-up
was constantly flushed with purified and dried nitrogen gas. Settled crystals
were washed and then freeze-dried and characterized using electron spectro-
scopy for chemical analysis (ESCA).
The stoichiometry of the samples was determined by measuring concen-
trations of dissolved species in the mineral supernatants. In these tests,
338
the conditioning of the minerals was done in a nitrogen glove box to mini-
mize the interference by atmospheric carbon dioxide. The molar ratios of
calcium/phosphorus (Ca/P) and calcium/carbonate (Ca/Ct) were 1.59 and
0.93 for apatite and calcite, respectively, and their calculated thermo-
dynamic solubility products were 10-115 and 10-8)45. ESCA was used to
determine the surface chemical composition of the freshly prepared (un-
treated) mineral surfaces. Comparison of the ESCA spectra with standards
provided by the Physical Electronics Division of Perkin-Elmer [31] showed
the samples to be pure calcite and apatite. Surface areas of the calcite and
apatite samples as determined by the BET technique using Quantasorb
were found to be 0.71 and 30.3 m2 g-l, respectively.
lnorganics
All inorganics such as KN~, KOH, K2C~, ~PO4, Ca(N~h, HN~ etc.
were of certified ACS grade purchased from Fisher Scientific Company.
Methods
Reactions at the mineraholution interface were studied in this work by
measuring the zeta potential as a function of pH under various conditions.
The zeta-potential values were measured by the electrophoretic technique
using the Lazer Zee Meter and the Zeta-Meter manufactured by Pen Ken Inc.
and Zeta-Meter Corp., respectively. The equilibration of the mineral for zeta-
potential measurements was done by stirring for 30 min 0.1 g of the mineral
with 100 ml of triply distilled water containing the required amount of KOH
or HNO3. Thirty minutes conditioning time was selected on the basis of
kinetic studies. Depending on the initial pH, significant changes in pH and
zeta potential were observed only within the first 10-20 min of conditioning.
Furthermore, the data in the literature indicate that conditioning times for
flotation tests do not often exceed 10 min [6]. It is to be noted, however,
that several hours of conditioning are required for complete equilibrium [32,
34]. The equilibration as well as electrokinetic measurements were conducted
in the presence of atmospheric carbon dioxide.
Zeta-potential measurements were conducted in mineral supernatants
as well as in water. The supernatant was prepared by stirring the mineral
with water at the desired pH for 30 min followed by centrifugation. The
supernatant obtained was checked under an optical microscope to ensure
the absence of solid particles. The pH was then readjusted and the solutions
used for conditioning of the mineral prior to zeta-potential measurements.
To check for possible effects of variations in ionic strength, experiments
were conducted also in 2 X 10-3 kmol m-J KN~.
Additional measurements were made in mixed supernatants prepared
by combining 50 ml each of the supernatants of calcite and apatite pre-
pared at about the same pH.
l
~
<j(
J:
0
..
I
;,
0
~
fa
1
oW
r
0
.-4
X
c-.
--
.0
--
~
i
.-
--
.5
~
i
.
'0
;
.s
'u
i
0
";
.;;
!
8.
.e
N
.-4
V
~
339
340
X
Q.
~
Q.
o~
~
r
S
i
~
r
Q
X
~
--
e.
'Ii
j
~
11;
i
.
0
..
-.g
$;
8-
~
N
Go
os:
d
0
d
.
!
'8.
='
.
Go
'u
..
u
...
0
~
r&1
~
.j
rz.
341
..
0
~
r
S
'0
S
~
i'
0
...
X
~
--
:e.
t
~
~
--
~
41
'u
"OJ
u
0
:i
=
41
&.
~
41
N
41
~
=
0
~
=
~
!
~
~
.
GI
~
&
~
0
.
U
GI
rz1
C')
~
~
X
Q.
x
~
343
:J:
Q.
..;
c
,.,
a
~
~
~
GI
~
~
Co
.
..s
3
i
.
--
~
i
i
'41
Co
~
~
GI
~
'u
g
.S
GI
~
'(3
g
~
~
0
:3
~
C
QI
i
.
~
N
aD
~
rz
344
In order to identify clearly the effects of supernatants, the results in
Figs 2(b) and 3(b) (experiments under controlled ionic strenth conditions)
are replotted in Fig. 4. The results show calcite and apatite to interchange
approximately their isoelectric points in the supernatants of each other.
The observation of Ananthapadmanabhan and Somasund~, reported
in Ref. [6], for the effect of the supernatants of calcite and apat.ite on
their flotation behavior must be noted here. These effects are obvi~,,-
due to the dissolved mineral species present in the supernatants. It )s-iD-
teresting to note that the flotation of these minerals was found t9---be af-
fected also by their own supernatants. However, the zeta potehtial of a
mineral is not expected to be affected to any appreciable extent by its own
supernatant if similar conditions are maintained since the equilibrium
concentration of the potential determining ions on the surface should
be identical whether treated in water or in their own supernatants. Figure
5 showing the results of the zeta potential of calcite and apatite in their
own supernatants indicates that, indeed, the zeta potentials are only slightly
affected.
M5
The results of the zeta-potential measurements in mixed supernatants,
presented in Fig. 6, further show the effects of dissolved species; if calcite
and apatite are present as a 1:1 mixture (weight basis), the two minerals
have identical surface-charge characteristics in the basic pH region.
Evidently, these effects are due to some or all of the dissolved mineral
species. In order to identify the species actually responsible, zeta-potential
measurements of calcite and apatite were conducted in inorganic solutions
containing each species. The results obtained in the presence of Ca(N~h,
K2C~ and K2HPO4 are given in Figs 7 to 9. Data given in Fig. 7 for the
zeta potential of apatite in Ca(N~h solutions of several concentrations
show that the surface of the mineral is more positively charged in Ca(N~h
solutions at all pH and levels of concentration tested. This is in agreement
with the results in the literature [4, 30, 34] .
The effect of K2C~ on the zeta potential of apatite (Fig. 8) is found
to be minimal in the concentration range studied in accord with the fact
that the carbonate will be present mostly in the HCO; form in the pH
range where apatite is negatively charged and, therefore, cannot be expected
60
40
>
E
~
-J
~
..-
Z
t&J
to
Q.
<r
..-
t&J
N
-20
-40
5 6 7 8
9
pH
10
12
13
Fig. 7. Effect of Ca(NO,h addition on the zeta potential of apatite.
20
0
346
1l
I
.
0
:i
...
c
.8
8-
i
~
oS
c
0
c
0
"D
:a
'Q
.
o'
~
0
]
rz1
~
.~
~
lAW) lVIJ.N3J.Od VJ.3Z
!i.
1
.
0
:j
j
8.
.s
=
CI
A
...
Q
0
Q
0
:.;
'0
'0
.
0
0
~
0
!
rz1
~
.'w
~
to produce any specific effects. On the other hand, high concentrations of
K1~ can be expected to cause the precipitation of CaC~ on the surface
of apatite to produce more positive zeta-potential values. However, relatively
higher concentrations of carbonate are required for the calcite precipitation.
While the concentrations of K1~ u~ in the experiments (10-3 and 10-1
kmol m-3 K1C~) would be insufficient for this process, calcite supernatants
could contain enough carbonate to effect the precipitation. The possibility
of surface conversion processes from a thermodynamic point of view is
discussed Jater in this section.
The effect of K3PO4 addition shown in Fig. 9 indicates that the surface
of apatite has become more negatively charged in the pH range examined
in agreement wiUl past results [27, 35]. This decrease is mainly due to the
role of phosphate as potential determining ions for apatite.
Figures 10 and 11 present results for the zeta potential of calcite in
the above inorganic solutions. As can be seen from Fig. 10, the effect of
Ca(NO3)1 addition is to make the surface of calcite more positively charged.
In the concentration range tested, only positive potentials were recorded
over the entire pH range.
60[
40~
~
CALCITE
kmoVm3 Co (~)2
"
"
,
--'\.--
\
WATER '-
\
'\
\
\
-4oL
5
\
\
.L
12
-l.
6
I ~,
1 .
.L
9
I I
10 II 13
pH
Fig. 10. Effect of Ca(NO.>' addition on the zeta potential of calcite.
348
'>
_E-
-'
ct
.-
Z
'"
.-
~
ct
.-
'"
N
pH
Fig. 11. Effect of K,PO. addition on the zeta potential of calcite.
Similar to the effects on apatite, K3PO4 addition reduces the zeta-potential
values of calcite (Fig. 11). It must be noted that zeta-potential values ob-
tained for calcite in 10-3 kmol m-3 K3PO4 closely agree with those obtained
for apatite in water. As shown later, even less than 10-9 kmol m-3 total
phosphate can cause precipitation of apatite from calcite supernatants under
neutral pH conditions. The observed changes in zeta potential can there-
fore be attributed to the precipitation of apatite on the calcite surface.
Analysis of the zeta-potential results presented above show the observed
shifts in the isoelectric point of calcite and apatite in the supernatants of
each other to be the result of many complex surface processes. The mecha-
nism of surface charge generation for these minerals has been examined
in detail during an earlier investigation [4, 5, 34, 35]. The electrokinetic
behavior of apatite and calcite in various inorganic solutions has also been
determined [32, 35]. These properties, however, have not been studied
for heterogeneous mineral systems. In a recent investigation [36] we have
shown from theoretical considerations that, depending on the solution
conditions, apatite surface can be converted to calcite, and vice versa
through surface reaction or bulk precipitation of the more stable phase.
349
The observed changes in the electrokinetic properties of calcite and apatite
can therefore be examined on the basis of the mineral-solution chemical
equilibria involving dissolved mineral species.
From studies of solubility isotherms for apatite and calcite at 2SO C
[19,23,37], the singular point for these minerals is identified to be 9.3.
Above this pH calcite is more stable than apatite. The implication is that,
if apatite is brought in contact with alkaline solutions (pH> 9.3) previously
equilibrated with calcite, calcite can precipitate on the surface of apatite.
From similar considerations, the conversion of calcite to apatite is also pos-
sible below the singular point. We have shown theoretically [36] that, if
calcite supernatants are prepared above pH 9.3, the total carbonate present
in these solutions can exceed the amount required for the conversion of
apatite to calcite as illustrated in Fig. 12(a). As shown in Fig. 12(b), the
total phosphate in equilibrium with apatite below the singular point can
also convert calcite surface to apatite under appropriate conditions. As
mentioned earlier, much lower phosphate concentrations are required in
this case. Our investigation on the solubility behavior of these minerals
based on supernatant analysis under various solution conditions [38] has
4 I
Q
2
HAP - CALCITE
If)
E
"
"0
E
~
-
!oJ
I-
oct
Z
0
m
~
oct
U
-J
oct
I-
0
I-
...
0
0
-J
(OPEN)
., ""'"
~
/
~
, /
~~~~~
(OPEN)
(CLOSED)
~tot/CALCITE-
-10
-12
4
6
8
10
12 14
pH
Fig. 12(a). Illustration of the conversion of apatite to calcite.
0
-2
-4
-6
-8
350
4
2
~
E
"0
E
oK
--
I&J
'4
x
Q..
(/)
0
X.
Q..
-J
C
6
t-
~
Co?
0
-J
10
-12
pH
Fig. 12 (b). illustration of the conversion of calcite to apatite [36].
also shown that, indeed, the precipitation of these minerals is quite pos-
sible.
The surface conversion of calcite and apatite is further illustrated by
the results of recent spectroscopic investigations in our laboratory using
ESCA [39]. The minerals were conditioned under solution conditions
similar to those used for zeta-potential studies. The data given in Fig. 13
show that, when apatite is conditioned in the supernatant of calcite at
pH ,..., 12, its surface exhibits spectroscopic properties characteristic of
both calcite and apatite. This behavior is attributed to the precipitation
fo calcite on apatite.
On the basis of the electrokinetic, spectroscopic, flotation, dissolution
and precipitation behavior of these minerals in heterogeneous systems,
as well as the analysis of the mineral-solution chemical equilibria govern-
ing them, conditions for selective flotation have been discussed in detail
in a recent publication [40].
0
-2
-4
-6
-8
351
-
cn
I&J
0
0
a:
I-
<.J
I&J
-J
I&J
11-
0
a:
I&J
m
~
~
z
>-
I-
cn
Z
I&J
I-
Z
H
'--"'-"'\ 1
CALCITE IN WATER
APATITE IN CALCITE SUP.
295 290 285 280 275
81 NO ING ENERGY, eV
Fig. 13. ESCA spectra of C(1s) peak of apatite conditioned in calcite supernatant at
pH- 12 [39).
SUMMARY AND CONCLUSIONS
Dissolved mineral species playa major role in the interfacial behavior
of sparingly soluble minerals such as calcite and apatite. Electrokinetic
studies were conducted on the calcite--apatite heterogeneous system to
develop an understanding of the manner in which dissolved species affect
this behavior. The isoelectric points of calcite and apatite in both water
and KNO3 solutions are 10.5 and 7.4, respectively. When the minerals
are conditioned in the supernatants of each other, they approximately
352
interchange their points of zero charge; when apatite is conditioned in the
supernatants of calcite, its isoelectric point shifts from 10.5 to about 6.5
(8.0 in the presence of 2 X 10-3 kmol m-3 KN~) towards that of apatite.
Similarly, the isoelectric point of apatite is shifted from 7.4 to about 11.0
towards that of calcite. If calcite and apatite are present as a 1: 1 mixture
(weight basis), the two minerals have identical surface-charge characteristics
in the basic pH range. These effects are discussed on the basis of mineral-
solution equilibria controlling these systems. Thermodynamic considerations
indicate that the apatite surface can be converted to calcite in the super-
natant of the latter and vice cersa. Based on the results of the electrokinetic,
spectroscopic, flotation, dissolution and precipitation studies, the observed
shifts in isoelectric points are attributed to the precipitation of one mineral
on the surface of the other.
ACKNOWLEDGEMENTS
Support of the National Science Foundation (CPE-83-04059) and Florida
Institute of Phosphate Research is greatly acknowledged. We wish to thank
K.P. Ananthapadmanabhan and Kenneth Wong for helpful discussions.
REFERENCES
1 E. Sorensen, J. Colloid Interface Sci., 45 (1973) 601-607.
2 J.E. Lawver, R.L. Wiegel, R.E. Snow and C.L. Huang, Benefication of Dolomitic
Florida Phosphate Reserves, XIVth Int. Miner. Proc. Congress, Session IV, Flotation,
Paper IV (20), CIMM, Ontario, Canada, 1982.
3 P. Somasundaran, On the Problem of Separation of Calcite From Calcareous Apatite,
XIth Int. Miner. Proc. Congress, Instituto di arte mineraria e preparazione dei minerali,
Cagliari, Vol. 2, 1975, pp. 155-156.
4 P. Somasundaran and Y.W. Wang, Surface Chemical and Adsorption Properties of
Apatite, in D.N. Mishra (Ed.), Adsorption and Surface Chemistry of Hydroxyapatite
Plenum, New York, NY, 1984, pp. 129-149.
5 H.S. Hanna and P. Somasandaran, Flotation of Salt-Type Minerals, in M.C. Fuerstenau
(Ed.), Flotation, A.M. Gaudin Memorial Vol. 1, AIME, New York, NY, 1976, pp.
197-271.
6 K.P. Anathapadmanabhan and P. Somasundaran, The Role of Dissolved Mineral
Species in Calcite-Apatite Flotation, Minerals and Metallurgical Processing, SME
AIME, 1 (1984) 36-42.
7 T_W. Healy, personal communications, 1981.
8 M.C- Fuerstenau and D.A. Eigillani, Trans. AIME, 235 (1966) 405.
9 F.F. Aplan and D. W. Fuerstenau, Principles of Non-metallic Flotation, in D. W.
Fuerstenau (Ed.), Froth Flotation, 50th Ann. Vol., AIME, New York, NY, 1962,
p.170.
10 R.W. Smith and S. Akhtar, Cationic Flotation of Oxides and Silicates, in D.W.
Fuerstenau (Ed.), Flotation, A.M. Gaudin Memorial Vol. I, AIME, New York, NY,
1976, p. 87.
11 J.B. Clemmer, B.H. Clemmons, C. Rampacek, M.F. Williams and R.H. Stacey, U.S.
Bur. Mines Rep. Invest., 3799, 1945.
12 A_M. Gaudin and A. Rizo-patron, Trans. AIME, 153 (1942) 462.
353
13 S.R.B. Cooke and M. Digre, Trans. AIME, 184 (1949) 299.
14 H.J. Modi and D.W. Fuerstenau, Trans. AIME, 217 (1960) 381.
15 M.C. Fuerstenau and J.D. Miller, Trans. AIME, 238 (1967) 153-160.
16 G.W. Wolstenholme and J.H. Schulman, Trans. Faraday Soc., 46 (1950) 475.
17 I. Iwasaki and H. Soto, Fundamentals of Phosphate Flotation, Prog. Rep., (1981) 22.
18 W. Stumm and J.J. Morgan, Aquatic Chemistry, 2nd edn, Wiley-Interscience, New
York, NY, 1981.
19 W.E. Brown and L.C. Chow, Chemical Properties of Bone Mineral, Annu. Rev. Mater.
Sci., 6 (1976) 213-236.
20 E.C. Moreno, T .M. Gregory and W.E. Brown, Preparation and the Solubility of
Hydroxyapatite, J. Res. Nat. Bur. Stand Sect. A, 72 (1968) 773-782.
21 F. Morel, Principles of Aquatic Chemistry, Wiley-Interscience, New York, NY, 1983.
22 G.A. Parks, Am. Mineral., 57 (1972) 1173.
23 W.E. Brown, Physicochemistry of Apatite Dissolution, Physicochimie et Crista1logra-
phie des Apatites d'Interet Biologique: Colloques Internationaux C.N.R.S., No. 230,
1973, pp. 355-368.
24 S. Polcn, Interaction Between Reagents and Minerals during Flotation of SnO."
Fe.O], and CaCO], Izv. Vyssh. Uchebn. Zaved. Chern Metal., 1 (1958) 11-23.
25 S.C. Sun, R.E. Snow and W.I. Purcell, Characteristics of Florida Leached Zone Phos-
phate Ore With Fatty Acids, Trans. AIME, 208 (1957) 70.
26 M.A. Eigeles, Selective Flotation of Non-8ulfide Minerals, Progress in Mineral Dressing,
Trans. 4th Int. Mineral Dressing Congress, Stockholm, Almqvist and Wiksell, 1958,
pp.591-609.
27 J.M. W. MacKenzie and S.K. Mishra, Zeta-Potential Studies on Apatite, Paper presented
at the 4th National Convention, Australian Chemical Institute, Canberra, Australia,
1970, p. 43.
28 J.C. Le Bell and L. Lindstrom, Electrophoretic Characterization of Some Calcium
Minerals, Finn. Chem. Lett., (1982) 134-138.
29 J. V. Calara and J.D. Miller, The Electrokinetic Behavior of Fluorite in the Absence
of Surface Carbonation, Colloids Surfaces, submitted.
30 Y.H. Wang, Zeta-Potential Studies on Hydroxyapatite, M.S. Thesis, Columbia Univer-
sity, New York, NY, 1975.
31 Handbook of Photoelectron Spectroscopy for Chemical Analysis, Physical Electronics
Division, Perkin-Elmer, Minnesota, U.S.A., 1978.
32 P. Somasundaran and G .E. Agar, Zero Point of Charge of Calcite, J. Colloid Interface
Sci., 24 (1967) 433-440.
33 R.K. Mishra, S. Chander and D. W. Fuerstenau, Effect of Ionic Surfactants on the
Electrophoretic Mobility of Hydroxyapatite, Colloids Surfaces, 1 (1980) 105.
34 P. Somasundaran, Zeta Potential of Apatite in Aqueous Solutions and Its Change
During Equilibration, J. Colloid Interface Sci., 27 (1968) 659.
35 P. Somasundaran and G.E. Agar, Further Streaming Potential Studies of Hydroxy-
apatite in Inorganic Solutions, Trans. AIME, 252 (1972) 348-352.
36 P. Somasundaran, J.O. Amankonah and K.P. Ananthapadmanabhan, Mineral-
Solution Equilibria in Sparingly Soluble Mineral Systems, Colloids Surfaces, 15 (1985)
309-333.
37 S.K. Mishra, The Electrokinetics of Apatite and Calcite in Inorganic Electrolyte
Environment, Int. J. Miner. Process., 5 (1978) 69-83.
38 J.O. Amankonah, P. Somasundaran and K.P. Ananthapadmanabhan, Effect of Dis-
solved Mineral Species on the Dissolution/Precipitation Characteristics of Calcite and
Apatite, Colloids Surfaces, 15 (1985) 295-307.
39 P. Somasundaran, J.O. Amankonah and K.P. Ananthapadmanabhan, Spectroscopic
Analysis of Calcite-Apatite Interactions Using ESCA, unpublished results.
40 P- Somasundaran, J.O. Amankonah and K.P. Ananthapadmanabhan, Calcite-Apatite
Interactions and Their Effects on Selective Flotation Using Oleate. Accepted for
presentation at the XV Int. Miner. Proc. Congress, Cannes, France, June 1985.