You are on page 1of 9

Minerals Engineering 79 (2015) 116124

Contents lists available at ScienceDirect

Minerals Engineering
journal homepage: www.elsevier.com/locate/mineng

Preparation of magnesium hydroxide from serpentinite by sulfuric acid


leaching for CO2 mineral carbonation
Qing Zhao a,b,, Cheng-jun Liu a, Mao-fa Jiang a, Henrik Saxn b, Ron Zevenhoven b
a
b

Key Laboratory for Ecological Metallurgy of Multimetallic Ores (Ministry of Education), Northeastern University, Shenyang 110819, China
Thermal and Flow Engineering Laboratory, bo Akademi University, bo/Turku 20500, Finland

a r t i c l e

i n f o

Article history:
Received 11 February 2015
Revised 30 May 2015
Accepted 1 June 2015

Keywords:
Magnesium hydroxide
Serpentinite
Sulfuric acid leaching
CO2 mineral carbonation

a b s t r a c t
Carbon capture and storage (CCS) by mineral carbonation is a promising way for CO2 emissions mitigation that has been under studied for decades. In this work, the preparation of magnesium hydroxide from
Finnish serpentinite using sulfuric acid leaching as the rst step of a CO2 mineral carbonation process was
studied. Some details of leaching behavior of the ore were revealed and a valuable metal was recovered in
this study. It was found that leaching yield of magnesium increased with sulfuric acid dosage, limited by a
product layer formed on the ore particles, resulting in incomplete serpentinite decomposition. Agitation
and ultrasonication were demonstrated to be effective in controlling the thickness of product layer. About
95% of iron was recovered from the leachate and leaching residues and valuable Fe-rich substances were
obtained as by-products. After the iron extraction, a ne Mg(OH)2-rich powder could be prepared from
the Mg-rich solution by precipitation using sodium hydroxide solution.
2015 Elsevier Ltd. All rights reserved.

1. Introduction
In recent decades, the greenhouse effect has become an issue of
common concern worldwide. Carbon capture and storage (CCS)
is considered as one of the main options for alleviating emissions
of CO2 from human activities. Ocean, geological and mineral
sequestrations are three candidates of large-scale technologies
for CCS. In ocean sequestration, CO2 could be directly injected into
the ocean by moving ships, stationary points or by long,
bottom-mounted diffusers, but the detrimental effects on marine
ecosystems, especially in sites with high CO2 concentrations, cannot be ignored (Israelsson et al., 2010). Geological sequestration
is another option where CO2 can be stored below ground by injecting CO2 into deep geological formations. However, injection sites
require proper permeable geological formations thus limiting the
application area of this approach: Furthermore, comprehensive
geophysical investigations need to be carefully carried out before
geological sequestration (Zhang and Song, 2014). After plenty of
evaluations, scientists now believe that mineral sequestration is
one of the most promising method for CO2 sequestration to many
countries (Goff et al., 2000). Moreover, it is very unlikely that CO2
captured in a thermodynamically and geologically stable material
Corresponding author at: Mailbox 313, Northeastern University, Wenhua Road,
Heping District, Shenyang, China.
E-mail address: qingzhao0121@gmail.com (Q. Zhao).
http://dx.doi.org/10.1016/j.mineng.2015.06.002
0892-6875/ 2015 Elsevier Ltd. All rights reserved.

(Sanna et al., 2013) would be released to the atmosphere again,


making post-storage monitoring unnecessary.
Serpentinite, containing mainly serpentine, is an excellent feedstock material for CO2 mineral carbonation because of the large
quantities available, extensive distribution and relatively low hardness (Nduagu et al., 2012a). Magnesium hydroxide could be prepared from serpentine by various methods for subsequent CO2
sequestration conducted at high temperature (>500 C) and high
pressure (>20 bar) atmosphere (Romo et al., 2013). bo
Akademi University has been performing intensive research within
this topic for about a decade and is developing a process route
toward industrial application, in which magnesium in a rst step
is extracted from Mg-rich silicate via solidsolid reaction using
recyclable ammonium sulfate or ammonium bisulfate at temperatures of 400500 C (Nduagu et al., 2012b; Romo et al., 2013).
An acid leaching process is another option for magnesium
extraction at relative low temperature, in which the magnesium
ion could be leached out from the ore into solution, and the magnesium hydroxide can be easily prepared by adjusting pH value of
the leachate (Alexander et al., 2007). Sulfuric acid, hydrochloric
acid, nitric acid, formic acid and acetic acid have all been investigated as solvents and the work has showed sulfuric acid to be
the most effective reagent in serpentine dissolution (Teir et al.,
2007a,b). Some activation pretreatments prior to the leaching process were demonstrated to improve mineral conversion
(Maroto-Valer et al., 2005). Kim and Chung (2002) studied the

Q. Zhao et al. / Minerals Engineering 79 (2015) 116124

effect of mechanical preprocess on leaching efciency, reporting


that almost all magnesium and iron was leached into the solution
from serpentine in only 5 min after 240 min ball-milling pretreatment. However, the high energy requirement from this long grinding duration works against the original purpose of CO2 emissions
reduction. Moreover, the high cost in recycling the acid is still a
challenge for industrialization.
Kodama et al. (2008) conducted metallic elements extraction
experiments from silicates, reporting that a Si-rich phase formed
on the surface of the ore particles during the leaching process,
which signicantly limited the diffusion of reactive ions.
However, a detailed understanding of the role of this product layer
and ways to remove it is still lacking.
Researchers admitted that the economy of CO2 sequestration is
a key factor for future technology deployment (Olajire, 2013).
Besides reducing the high energy requirement by optimizing process conditions and parameters, it is possible to solve the problem
of process economics by producing valuable by-products in the
sequestration process. It has, for example, been demonstrated that
the utilization of various iron compounds from serpentine as raw
materials for the iron- and steel-making industry could be feasible
solution to offset CO2 sequestration costs (Nduagu et al., 2012c;
Romo et al., 2012). Solvent extraction by organic extractants is a
well-known approach for accomplishing the complex separations
of metals from multi-element acid solutions. Many organic extractants can be easily recovered and repeatedly used, controlling the
follow-up input of solvent extraction process. However, little work
has yet been reported on how to extract valuable metals from
serpentine leachate by solvent extraction.
This paper aims at a less energy-intensive process route to prepare magnesium hydroxide for CO2 sequestration from Finnish serpentinite. Sulfuric acid leaching and solvent extraction were
conducted for serpentinite decomposition and iron extraction.
The leaching behavior of metallic elements, the effects of agitation
and ultrasonication treatment on this passivating layer, and the
optimization of extraction conditions were investigated in the
current work.

117

Fig. 1. XRD pattern of Finnish Hitura serpentinite.

Serpentinite powder with a size smaller than 74 lm was


obtained by grinding and screening, which was smaller than the
previous study (Teir et al., 2007b). This was used for studying the
leaching behavior of metallic elements in the ore. A series of
tests using a serpentinite lump was also carried out to elucidate
the product layer formation and removal behavior in the
leaching process. Thus, the rock was cut and polished to get a
smooth at surface. A LEO 1530 Gemini with a scientic ultra
dry silicon drift detector was employed in scanning electron
microscopy-energy-dispersive X-ray spectroscopy (SEMEDS)
analysis. A photograph and a SEM image of the lump surface are
shown in Fig. 2. On the basis of results from EDS connected with
SEM, it could be conrmed that plenty of bright phases, with sizes
in the range 1030 lm, distributed in the dark silicate matrix were
magnetite. This implies that it would be difcult to effectively separate magnetite by magnetic separation from the Mg-rich silicate
since it has not been liberated. As a result, iron recovery tests both
for leachate and leaching residues were conducted by solvent
extraction and magnetic separation, respectively, after the sulfuric
acid leaching process.

2. Experimental
2.2. Methods

2.1. Materials
Serpentinite used in this study was taken from the stockpile
of the Finnish Hitura nickel mine of Belvedere Resources Ltd.
(formerly Outokumpu Mining Oy). Inductively coupled
plasma-optical emission spectrometry (ICP-OES) analysis was carried out by Varian Vista-MPX against suitably calibrated standards
(520 ppm) on aqueous extracts from 100 mg samples diluted to
500 ml to detect the chemical composition of the serpentinite;
the results are given in Table 1. The phase composition of serpentinite was analyzed by Philips Xpert X-ray diffraction (XRD) with
Cu Ka source (k = 1.5418 ) over the range 2h = 1570 at a step
size of 0.008 and specied by Crystallographica Search-Match
(CSM) software with the Powder Diffraction File (PDF) databases
from International Centre for Diffraction Data (ICDD), which can
be seen in Fig. 1. The ore was found to contain 83 wt.% serpentine
(Mg3Si2O5(OH)4). In the remaining impurities, magnetite (Fe3O4)
accounted for the largest fraction (82 wt.%). The Mg/Fe mass ratio
is 2.2, BET surface area 26.45 m2 g1 m2/g and pore volume
0.0347 cm3 g1.

Table 1
Chemical composition of Finnish serpentinite (wt.%).
Mg

Fe

Ca

Ni

Al

Cr

Cu

Si

21.80

10.10

0.34

0.28

0.02

0.01

0.08

11.60

2.2.1. Ore powder experiments


10 g of serpentinite was poured into an Erlenmeyer ask with a
certain amount (the ratios of ore powder mass and acid volume
were 0.1, 0.2, 0.4 and 1 g mL1) of sulfuric acid of certain molarity
(2, 3, 4 and 5 mol L1), and then agitation was started and maintained during the whole leaching process. After a certain period
of time (10, 20, 30, 40, 50, 60, 90 and 120 min), the leachate was
obtained by ltration and analyzed by ICP-OES. The leaching yield
of metallic elements, expressed as the mass ratio of metallic elements in the leachate and in the raw material, was determined.
The experimental set-up for serpentinite powder leaching is shown
in Fig. 3(a).
2.2.2. Ore lump experiments
To study the product layer formed on ore particles, a batch of
experiments was carried out using a serpentinite lump particle.
Sulfuric acid (200 mL 4 mol L1) was added into a glass beaker
and the lump was immersed into the solution with the smooth at
surface facing upward. Agitation and ultrasonication treatment
were employed in some tests in an attempt to remove the product
layer. After 2 h of leaching, the lump was taken out of the acid solution and was carefully washed with deionized water to avoiding
morphology changes, followed by a drying process at room temperature. The apparent morphology of the lump was investigated
using Olympus 3D measuring laser microscope (3DMLM).

118

Q. Zhao et al. / Minerals Engineering 79 (2015) 116124

Fig. 2. Photograph (left) and SEM image (right) of Finnish serpentinite lump.

Fig. 3. Schematic illustration of the experimental set-up for leaching the Finnish serpentinite. (a) Leaching Finnish serpentinite powder; (b) Leaching Finnish serpentinite
lump without treatment; (c) Leaching Finnish serpentinite lump with agitation; (d) Leaching Finnish serpentinite lump with ultrasonication. 1 Variable speed blender; 2
Retort stand; 3 Erlenmeyer ask; 4 Serpentinite powder; 5 Sulfuric acid; 6 Glass ask; 7 Water; 8 Serpentinite lump; 9 Ultrasonic vibration equipment.

Fig. 3(b)(d) present a schematic illustration of the equipment and


procedure.

phases was achieved. Reactions of P507 saponication and extraction of Fe3+ are as follows:

2.2.3. Solvent extraction


2-Ethylhexyl dihydrogen phosphate (P507) is regarded as an
excellent extractant for Fe3+ in chemical industry. A certain
amount of hydrogen peroxide was needed to oxidize all of the
Fe2+ ions in the leachate to Fe3+ ions before the solvent extraction.
After this, some industrial P507 (>95 wt.%) was mixed with sulfonated kerosene in different ratios to reduce the viscosity of
P507, and then some 10 wt.% sodium hydroxide solution was
added into the diluted extractant to saponify the P507. The saponied extractant and the serpentinite leachate were mixed in a separating funnel at room temperature, shaken for a few minutes to
extract Fe3+ into organic phase followed by a short time of standing, ensuring that a complete separation of organic and aqueous

Fe3+ and Mg2+ contents left in aqueous phase were determined


by ICP-OES to calculate the extraction yield according the mass
ratio of metallic ions in aqueous phase before and after solvent
extraction.
2.2.4. Product preparation
Magnesium hydroxide was precipitated from the aqueous
phase when the pH value reached 10 using sodium hydroxide
solution. 4 mol L1 Hydrochloric acid was employed to strip iron
from organic phase into aqueous acid solution for valuable
Fe-containing products preparation. After evaporation of the
Fe-rich solution and lter cake washing, iron oxide was obtained.
Almost all of the P507 was regenerated in the stripping process

Q. Zhao et al. / Minerals Engineering 79 (2015) 116124

119

Fig. 4. Flow sheet of the process route.

Fig. 5. Effect of sulfuric acid on leaching yield of magnesium.

for reutilization, and some HCl and vapor were also recovered in a
glass condenser pipe during the evaporation phase. Magnetite was
collected from the leaching residues by magnetic separation with a
magnetic eld intensity of 50 mT using a DTCXG-ZN50 magnetic
separator. A ow sheet of the overall process route proposed in this
study is given in Fig. 4.

4 mol L1 showed no further improvement on magnesium leaching. Therefore, a reasonable acid dosage was 4 mol L1 sulfuric acid
with a ore mass/acid volume ratio of 0.4 g mL1.
Leaching behavior of the Finnish serpentinite was investigated
using the optimal sulfuric dosage. A batch of tests was conducted
for different leaching durations, and the concentrations of all
metallic elements in the leachates were determined by ICP-OES.
The changes of leaching yield with duration are shown in Fig. 6,
in which the error bars was the average value of the results of
two parallel experiments. It was found that recovery rates of all
metallic elements steadily increased with duration until 30 min,
while no notable change occurred from 30 min to 120 min. The
optimal duration obtained in this study was shorter than it
reported by Alexander et al. (2007) who worked on the serpentine
from the Cedar Hills quarry in SE Pennsylvania using similar the
leaching conditions, demonstrating the high reactivity of the
Finnish serpentinite. Most of the metallic elements except iron
and nickel were leached out from the ore, with a leaching yield
for magnesium of about 86%. A leachate with about 3 mol L1 of
Mg2+ was obtained from this process, which could be used to prepare magnesium hydroxide for the CO2 capture after iron recovery.
Therefore, a leaching time of 30 min is considered long enough for
magnesium extraction in 4 mol L1 sulfuric acid.

3. Result and discussion


3.1. Leaching behavior of ore
Both concentration (2, 3, 4 and 5 mol L1) and volume (the
ratios of ore powder mass and acid volume were 0.1, 0.2, 0.4 and
1 g mL1) were simultaneously considered in 2 h-tests to study
the effect of sulfuric acid on extraction yield of magnesium, giving
the results shown in Fig. 5. The acidity and volume of sulfuric acid
proved to be signicant factors for the leaching yield of magnesium. When 2 mol L1 sulfuric acid was used in the leaching process, only about 70% of the magnesium could be extracted from
the serpentinite. Higher than 80% of leaching yield was achieved
in the experiments using acid with concentration higher than
3 mol L1. Furthermore, the acid volume could be reduced if high
acidity was employed, but sulfuric acid concentration exceeding

Fig. 6. Relationship between the leaching yield of metallic elements and the
leaching duration.

120

Q. Zhao et al. / Minerals Engineering 79 (2015) 116124

deposited on the surface of particles and acted as an obstacle


impacting the contact of the reactants. More details were investigated and are disclosed in the next section.
3.2. Product layer investigation

Fig. 7. XRD patterns of Finnish serpentinite and residues after different leaching
durations.

To clarify the reasons for the low leaching yield of iron and
nickel, the phase compositions of residues from 10 min, 30 min
and 50 min leaching tests were detected by XRD (Fig. 7) and
compared with the compositions of the original ore. The results
demonstrated the major phases of all samples to be Fe3O4 and
Mg3Si2O5(OH)4. The intensity of diffraction peaks of serpentine
weakened signicantly with leaching duration but did not show
any notable changes from 30 min to 50 min, coinciding with the
results obtained with ICP-OES. As for the diffraction peaks of magnetite, they decreased slower than the serpentine peaks because
magnetite barely dissolves in sulfuric acid solution of this acidity.
Therefore, some magnetite was left in residues after the leaching
process, leading a low leaching yield of iron. He (2010) proposed
that Ni-bearing serpentine is also a refractory phase in acid solution, so nickel may concentrate in residues after the leaching treatment. Experimental results of this study showed that the content
of nickel in the leaching residues reached 0.62%, which is considerably higher than the original 0.28%. This means that these residues
could be utilized as a secondary nickel resource in nickel extraction
industries. Kodama et al. (2008) suggested that the reason for the
incomplete extraction of magnesium is that a Si-rich phase

A polished serpentinite lump was immersed in 4 mol L1 sulfuric acid for 2 h, and was then removed and dried at room temperature or 90 C for 2 h. After this the apparent morphology was
studied using 3DMLM; some results are illustrated in Fig. 8.
Before the leaching treatment, a smooth ore lump surface (cf. 1
layer) can be seen in the gure. A glassy layer (cf. 2 layer) was
formed and covered the whole lump in room temperature tests.
By EDS and ICP-OES the elemental composition of this glassy phase
was detected to be silicon and oxygen (It should be noted that
neither EDS nor ICP-OES can detect hydrogen.). Additionally, no
diffraction peaks of new Si-bearing crystal appeared in any
patterns (cf. Fig. 7), demonstrating the amorphous form of this
Si-rich phase. A hump detected below 20 2-Theta cannot be
specied by any known standard, which was speculated to be a
function of the XRD itself. With reference to an Eh-pH diagram of
MgSiOH at 25 C and the glassy state (cf. 2 layer), this phase
generated in the serpentinite leaching should be the amorphous
silicic acid rather than silica. After a drying at 90 C, the product
layer was transformed into silica particles (cf. 4 layer) because
of the dehydration, explaining the confusion of Park et al. (2003)
and Kodama et al. (2008) on the phase determination of the
product layer. Moreover, when the leaching process is conducted
at high temperature like 160 C, silicate could transform into
amorphous silica directly (Zhao et al., 2014).
Furthermore, it was found that this product layer can be
easily removed, exposing a relatively smooth inner surface of
the lump (cf. 3 layer). Therefore, in the decomposition of serpentinite during the leaching process it could be speculated that
hydrogen ions diffused through this layer from the solution to
the surface of the ore particles to react with the inner core.
The shrinking process of the particles progressed in a uniform
way, so a smooth surface of the inner unreacted core was
obtained. There is no doubt that the removal treatment for this
passivating layer is critical for the leaching rate and for the
completion of the reactions.

Fig. 8. Apparent morphology of serpentinite lump before and after 2 h immersion in 4 mol L1 sulfuric acid.

Q. Zhao et al. / Minerals Engineering 79 (2015) 116124

121

Fig. 9. Apparent morphology of serpentinite lump leached with different treatments.

Some lump experiments with an agitation at 400 r min1 or


with an ultrasonication at the frequency of 40 kHz were conducted
to remove the product layer and compared with the immersion
test. The apparent morphology of all lumps after 2 h leaching in
4 mol L1 sulfuric acid was detected by 3DMLM, as shown in
Fig. 9. The serpentinite surface of the lump could be generally
observed through/below a glassy phase after the agitation test,
showing that the product layer was much thinner than the one
without any treatment. It can be inferred that the ow of solution
caused by stirring could effectively control the thickness of the
product layer in the leaching process, and the diffusion rate of
hydrogen ions through this layer, as a consequence, would not
signicantly decrease with duration. Therefore, the decomposition
of ore particles in sulfuric acid may be brought closer to completion. As for the ultrasonication test, a product-layer-free lump with
a clear surface was obtained. The explanation of this could be
attributed to the cavitation that occurs when the solution is subjected to rapid changes of pressure in ultrasonication treatment.
Cavities are formed where the pressure is relatively low, and the
voids implode when they are subjected to higher pressure, which
can generate an intense shockwave. The product layer cannot stay
attached to the surface of lump under the effect of this cavitation.
Furthermore, solution temperature was also elevated by above
50 C by the ultrasonication treatment, which favored the leaching
process as well. However, it is important to note that this treatment may have a signicant energy input requirement, which
should be considered in evaluating the appropriateness of the
process for overall emission reduction.
A set of batch ore powder tests with different treatments was
carried under the same conditions as the lump tests and some
results are presented in Fig. 10. The leaching yield of magnesium
is seen to improve from about 48% to 86% with the employment
of agitation, and exceeded 90% in an ultrasonication test. Iron
leaching was also elevated by these treatments. In addition to
the effect of a removal of the product layer discussed above, ore
particles could now uniformly distribute in the solution by stirring
and cavitation, ensuring a full exposure of the particles to the acid
solution. As a result of this, the decomposition of serpentinite was
more complete. Therefore, keeping in mind the energy input
requirements, the agitation treatment was concluded to be the
most viable approach to remove the product layer in powder tests
of this study.

Fig. 10. Leaching yield of magnesium in sulfuric acid with different leaching
methods.

3.3. Iron recovery


An industrial hydrogen peroxide with a 1.2 times volume than
the theoretical one was added in the leachate to oxidize Fe2+ to
Fe3+ ions for a subsequent solvent extraction using the organic
extractant P507. P507 is a widely used extractant in acid system
that exchanges with H+ in POH and becomes POFe in the
way of cation exchange mechanism (Wu et al., 2013).
The separation factor of Fe3+ and Mg2+ (bFe3 =Mg2 ) observed in
the current work is a signicant index for evaluation of the
separation result of Fe3+ and Mg2+, which dened by

bFe3 =Mg2

C Fe3 O =C Fe3 A
C Mg2 O =C Mg2 A

where C Fe3 O is the concentration of Fe3+ in the organic phase,


C Fe3 A is the concentration of Fe3+ in the aqueous phase, C Mg2 O
is the concentration of Mg2+ in the organic phase, and C Mg2 A is
the concentration of Mg2+ in the aqueous phase.
Some extraction parameters including pH value (1.00, 1.25,
1.50, 1.75 and 2.00), concentration of P507 (20, 30, 40 and

122

Q. Zhao et al. / Minerals Engineering 79 (2015) 116124

Table 2
Separation factor and extraction yield of Fe3+ and Mg2+ under different extraction conditions (wt.%).
No.

pH value

Concentration of P507 (vol.%)

Saponication rate (%)

Duration (min)

Fe3+%

Mg2+%

bFe3 =Mg2

1
2
3
4
5
6
7
8
9
10
11
12
13
14

1.00
1.25
1.50
1.75
2.00
1.50
1.50
1.50
1.50
1.50
1.50
1.50
1.50
1.50

40
40
40
40
40
20
30
50
40
40
40
40
40
40

40
40
40
40
40
40
40
40
0
20
60
40
40
40

6
6
6
6
6
6
6
6
6
6
6
2
4
8

47.3
76.7
98.1
95.2
94.7
55.8
83.1
99.1
77.0
87.9
97.8
92.2
98.2
98.9

4.0
3.2
6.3
5.2
7.1
1.2
2.9
13.7
3.1
4.4
9.0
5.8
6.1
6.4

22
108
768
361
252
125
158
608
112
183
490
175
768
768

Fig. 11. Effect of extraction conditions on extraction results.

50 vol.%), saponication rate (0%, 20%, 40% and 60%) and duration
(0, 2, 4, 6 and 8 min) were investigated with a constant phase ratio
of 1:1, dened as the ratio between organic volume and aqueous
volume. The separation factor and the extraction yield of both
Fe3+ and Mg2+ ions under different extraction conditions are report
in Table 2 and Fig. 11.
The relative separation of the curves gives the possibility for the
selective extraction of Fe3+ from Mg-rich solution. The experimental data reveals that the extraction for Fe3+ was higher than that for
Mg2+ in all tests. From the results with respect to pH, it can be concluded that the extraction yield of Fe3+ rose when pH increased
from 1.00 to 1.50, while it decreased slightly when the pH value
increased further. The extraction yield of Mg2+ did not exhibit a
notable change with pH in the range of 1.00 to 2.00. Therefore,

the highest separated factor of Fe3+ and Mg2+ appeared at


pH = 1.50 where about 98% of the Fe3+ was extracted from the
aqueous into the organic phase with a minor Mg2+ extraction of
about 6%. A batch of tests was implemented with different P507
dosages in a constant organic volume to investigate the effect of
extractant concentration on the extraction results. It was found
that the extractions for Fe3+ and Mg2+ were both improved with
an increase in concentration of P507 in the range 2060 vol.% while
40 vol.% was the optimal value for Fe3+ and Mg2+ separation in the
current study. Higher extractant concentration than the optimal
led to a better Mg2+ extraction but showed no signicant increase
in Fe3+ recovery. Furthermore, the separation efciency strongly
depends on the viscosity of extractant, and an emulsication
may occur when the viscosity rises too high. Therefore, 40 vol.%

Q. Zhao et al. / Minerals Engineering 79 (2015) 116124

is here considered an appropriate extractant concentration in this


work. Fig. 11 also presents the relationship between saponication
rate of P507 and extraction results, showing the maximum values
of extraction yield of Fe3+ and separation factor of the two metallic
elements in the tests at the point where the saponication rate of
P507 was 40%. The Fe3+ recovery did not experience any signicant
changes when more P507 was saponied. Moreover, the pH value
of the aqueous phase rose from its initial value of 1.50 to 3.39
when 40 vol.% P507 was employed. This demonstrates that more
hydroxide ions were released from hydrolyzate of Fe3+ than hydrogen ions provided by unsaponied P507, so some hydroxide ions
left in the aqueous phase led a decrease in the pH value. The data
obtained from duration tests can be also seen in Fig. 11. The results
showed that the extraction process of P507 for Fe3+ (and Mg2+) is
very fast: more than 98% of the Fe3+ could be extracted into organic
phase by saponied P507 within 4 min. The indices for the solvent
extraction treatment changed little when extraction duration
exceeded 4 min.
Hydrochloric acid was used as the stripping agent as its stripping capacity is better than that of sulfuric acid and nitric acid
according to preliminary experiments. 4 mol/L hydrochloric acid
was used with a phase ratio of 1:1 in a stripping process to recover
Fe3+ from the organic phase for subsequent preparation of
Fe-containing inorganic products. The experimental procedure for
solvent extraction was reported in Section 2.2.3. This treatment
was conducted repeatedly for 8 min, each time followed by the
ICP-OES analysis for Fe3+. The ndings show that about 62% of
Fe3+ was stripped into aqueous phase after one stripping period,
and 99% of Fe3+ was recovered from organic phase after four
stripping periods. Hardly any Mg2+ was stripped out by the
hydrochloric acid in the tests, so a Fe-rich solution was obtained.
When the iron recovery process is applied in industrial scale, it
is important to evaluate the possibility of material cyclic utilization. In this study, P507 was regenerated by the stripping treatment because H+ ions replaced the Fe3+ ions in Fe-bearing P507,
and the P507 thus recycled in this stage can serve as an extractant
again. HCl was also recovered in this study by evaporating the ferric chloride acid solution for hydrochloric acid solution regeneration, and then an iron oxide (>99 wt.%) was obtained in residues
as a by-product. The recycle of extractant and hydrochloric acid
gives this iron extraction method a high competitiveness in the
perspective of economics. But from an environmental point of
view, manufacturing ferric chloride by concentrating the Fe-rich
solution is a promising option for utilization of the iron extracted
from Finnish serpentinite without any HCl gas generation.
To recover the magnetite in the leaching residues, some magnetic separation tests with a magnetic eld intensity of 50 mT were
conducted and the same magnetic separation was also carried out
to ore powder for comparison. The magnetic separation yield was
found to be about 46% in the ore tests, while about 93% of iron was
recovered from leaching residues. Moreover, the content of Fe3O4
in magnetite concentrate was increased from 73.2% to 92.1% after
the leaching process. The reason for these improvements is that
plenty of magnetite aked from serpentine during the leaching
process by acid corrosion and solution stirring, so a higher recovery
rate and purity can be obtained due to a very high phase
separation.
3.4. Products
After the iron extraction, magnesium hydroxide in micron scale
was precipitated from the Mg-rich aqueous phase by a pH adjustment until the pH value reached about 10 (using sodium hydroxide
solution), followed by a drying for 2 h at 110 C. Magnesium
hydroxide and iron oxide prepared in this study were weighted
and detected by XRD and ICP-OES analysis, which, as expect, shows

123

the major phases to be Mg(OH)2 and Fe2O3 with few impurities.


The results also indicated that the recovery rate of magnesium of
the whole process, expressed as mass ratio between magnesium
in the magnesium hydroxide and in the original serpentinite ore,
was 83.7%. Moreover, solvent extraction and magnetic separation
recovered 95.3% of iron, so iron oxide and magnetite were concentrated as by-products. The iron oxide purity was higher than 99%,
and the content of Fe3O4 in the magnetite concentrate was about
92%, as reported in Section 3.3 of this study.
4. Conclusions
The preparation of magnesium hydroxide from Finnish serpentinite by sulfuric acid leaching for CO2 mineral carbonation was
investigated at ambient temperature and pressure. About 86% of
magnesium was leached out in 30 min by 4 mol L1 sulfuric acid
with the ore mass/acid volume of 0.4 g mL1, while some
magnetite and Ni-bearing serpentine was left in the leaching residues. Agitation and ultrasonication treatments were demonstrated
as effective methods for passivation or removal of product layer,
leading to a more complete decomposition of the ore. About 98%
of the iron was extracted in 4 min from the leachate by 40 vol.%
P507 with a saponication rate of 40%. Almost all of the iron was
stripped from organic phase by 4 mol L1 hydrochloric acid.
About 93% of the iron from leaching residues by magnetic separation, and a magnetite concentrate was obtained as a by-product.
Mg(OH)2-rich powder was nally prepared for CO2 mineral
carbonation. Recovery rates of magnesium and iron of the whole
process were 83.7% and 95.3%, respectively.
Based on the conclusions of this work, future research will focus
on a comprehensive utilization of other valuable metallic elements,
especially the nickel in the leaching residues, in a cleaner and
economical way. Another line of future work would be to reduce
acid consumption with the nal goal to implement this
low-energy process in industrial CO2 mineral carbonation.
Acknowledgements
The authors gratefully acknowledge supports by China
Scholarship Council (CSC) for the visit of Qing Zhao to bo
Akademi University, Finland. Program for New Century Excellent
Talents in University of Ministry of Education of China (No.
NCET-11-0077) and the Fundamental Research Funds for the
Central Universities (No. 130402020) are also acknowledged.
Funding from TEKES and Finnish metals industry within the SIMP
research program under the Finnish Metals and Engineering
Competence Cluster (FIMECC Oy) is gratefully acknowledged.
References
Alexander, G., Maroto-Valer, M.M., Gafarova-Aksoy, P., 2007. Evaluation of reaction
variables in the dissolution of serpentine for mineral carbonation. Fuel 86, 273
281.
Goff, F., Guthrie, G., Lipin, B., Fite, M., Chipera, S., Counce, D., Kluk, E., Ziock, H., 2000.
Evaluation of Ultramac Deposits in the Eastern United States and Puerto Rico
as Sources of Magnesium for Carbon Dioxide Sequestration. Los Alamos
National Laboratory and U.S. Geological Survey, New Mexico.
He, Z.X., 2010. Study on the comprehensive utilization of Ni-bearing serpentine.
MSc. Thesis. Central South University, Changsha (in Chinese).
Israelsson, P.H., Chow, A.C., Adams, E.E., 2010. An updated assessment of the acute
impacts of ocean carbon sequestration by direct injection. Int. J. Greenh. Gas.
Con. 4, 262271.
Kim, D.J., Chung, H.S., 2002. Effect of grinding on the structure and chemical
extraction of metals from serpentine. Part. Sci. Technol. 20, 159168.
Kodama, S., Nishimoto, T., Yamamoto, N., Yogo, K., Yamada, K., 2008. Development
of a new pH-swing CO2 mineral carbonation process with a recyclable reaction
solution. Energy 33, 776784.
Maroto-Valer, M.M., Fauth, D.J., Kuchta, M.E., Zhang, Y., Andresen, J.M., 2005.
Activation of magnesium rich minerals as carbonation feedstock materials for
CO2 sequestration. Fuel Process. Technol. 86, 16271645.

124

Q. Zhao et al. / Minerals Engineering 79 (2015) 116124

Nduagu, E., Bergerson, J., Zevenhoven, R., 2012a. Life cycle assessment of CO2
sequestration in magnesium silicate rock-A comparative study. Energy Convers.
Manage. 55, 116126.
Nduagu, E., Bjrklf, T., Fagerlund, J., Mkil, E., Salonen, J., Geerlings, H.,
Zevenhoven, R., 2012b. Production of magnesium hydroxide from magnesium
silicate for the purpose of CO2 mineral carbonation-Part 2: Mg extraction
modeling and application to different Mg silicate rocks. Miner. Eng. 30, 8794.
Nduagu, E., Fagerlund, J., Zevenhoven, R., 2012c. Contribution of iron to the
energetics of CO2 sequestration in Mg-silicates-based rock. Energy Convers.
Manage. 55, 178186.
Olajire, A.A., 2013. A review of mineral carbonation technology in sequestration of
CO2. J. Petrol. Sci. Eng. 109, 364392.
Park, A.A., Jadhav, R., Fan, L.S., 2003. CO2 mineral sequestration: chemically
enhanced aqueous carbonation of serpentine. Can. J. Chem. Eng. 81, 885890.
Romo, I.S., Gando-Ferreira, L.M., Zevenhoven, R., 2013. Combined extraction of
metals and production of Mg(OH)2 for CO2 sequestration from nickel mine ore
and overburden. Miner. Eng. 53, 167170.
Romo, I.S., Nduagu, E., Fagerlund, J., Gando-Ferreira, L.M., Zevenhoven, R., 2012.
CO2 xation using magnesium silicate minerals. Part 2: Process energy

efciency and integration with iron-and steelmaking. Energy Int. J. 41 (1),


203211.
Sanna, A., Wang, X., Lacinska, A., Styles, M., Paulson, T., Maroto-Valer, M.M., 2013.
Enhancing Mg extraction from lizardite-rich serpentine for CO2 mineral
sequestration. Miner. Eng. 49, 135144.
Teir, S., Kuusik, R., Fogelholm, C.J., Zevenhoven, R., 2007a. Production of magnesium
carbonates from serpentinite for long-term storage of CO2. Int. J. Miner. Process.
85 (13), 115.
Teir, S., Revitzer, H., Eloneva, S., Fogelholm, C.J., Zevenhoven, R., 2007b. Dissolution
of natural serpentinite in mineral and organic acids. Int. J. Miner. Process. 83 (1
2), 3646.
Wu, W.Y., Zang, F.Y., Bian, X., Xue, S.F., Yin, S.H., Zheng, Q., 2013. Effect of loaded organic
phase containing mixtures of silicon and aluminum, single iron on extraction of
lanthanum in saponication P507-HCl system. J. Rare Earth 31 (7), 722726.
Zhang, D.X., Song, J., 2014. Mechanisms for geological carbon sequestration. Proc.
IUTAM 10, 319327.
Zhao, Q., Liu, C.J., Shi, P.Y., Zhang, B., Jiang, M.F., Zhang, Q.S., Saxn, H., Zevenhoven,
R., 2014. Sulfuric acid leaching of South African chromite. Part 1: Study on
leaching behavior. Int. J. Miner. Process. 130, 95101.

You might also like