Synthesis of High Purity Precipitated CA

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Waste Management xxx (2015) xxx–xxx

Contents lists available at ScienceDirect

Waste Management
journal homepage: www.elsevier.com/locate/wasman

Synthesis of high-purity precipitated calcium carbonate during the


process of recovery of elemental sulphur from gypsum waste
M. de Beer a,d,⇑, F.J. Doucet b,⇑, J.P. Maree c, L. Liebenberg d
a
DST/CSIR National Centre for Nanostructured Materials, Council for Scientific and Industrial Research, PO Box 395, Pretoria 0001, South Africa
b
Council for Geoscience, Private Bag X112, Pretoria 0001, South Africa
c
Department of Environmental, Water and Earth Science, Faculty of Science, Tshwane University of Technology, Private Bag X680, Pretoria 0001, South Africa
d
Centre for Research and Continued Engineering Development (Pretoria), North-West University, South Africa

a r t i c l e i n f o a b s t r a c t

Article history: We recently showed that the production of elemental sulphur and calcium carbonate (CaCO3) from gyp-
Received 18 May 2015 sum waste by thermally reducing the waste into calcium sulphide (CaS) followed by its direct aqueous
Revised 13 August 2015 carbonation yielded low-grade carbonate products (i.e. <90 mass% as CaCO3). In this study, we used
Accepted 17 August 2015
the insight gained from our previous work and developed an indirect aqueous CaS carbonation process
Available online xxxx
for the production of high-grade CaCO3 (i.e. >99 mass% as CaCO3) or precipitated calcium carbonate
(PCC). The process used an acid gas (H2S) to improve the aqueous dissolution of CaS, which is otherwise
Keywords:
poorly soluble. The carbonate product was primarily calcite (99.5%) with traces of quartz (0.5%). Calcite
Gypsum waste
Indirect carbonation
was the only CaCO3 polymorph obtained; no vaterite or aragonite was detected. The product was made
Precipitated calcium carbonate up of micron-size particles, which were further characterised by XRD, TGA, SEM, BET and true density.
Calcium sulphide Results showed that about 0.37 ton of high-grade PCC can be produced from 1.0 ton of gypsum waste,
Valorisation and generates about 0.19 ton of residue, a reduction of 80% from original waste gypsum mass to mass
of residue that needs to be discarded off. The use of gypsum waste as primary material in replacement
of mined limestone for the production of PPC could alleviate waste disposal problems, along with con-
verting significant volumes of waste materials into marketable commodities.
Ó 2015 Published by Elsevier Ltd.

1. Introduction leaching of saline water into surface and underground water and
the generation of airborne dust (Mbhele et al., 2009). Not only
Disposal of industrial by-products is becoming an increasingly are these gypsum stacks unsightly, but they also occupy large areas
serious constraint on many industries. Gypsum waste of land and require long-term expenditures for maintenance and
(CaSO42H2O) is a calcium-rich solid residue generated in the min- monitoring (Tayibi et al., 2009). In the context of reducing the envi-
ing industry as a result of the treatment of acid mine drainage ronmental burden and enhancing economic benefit, technologies
(AMD). Acidic water is naturally generated by the weathering of for converting waste materials into products of commercial value
sulphide minerals that are exposed to atmospheric conditions dur- are in great demand. To achieve a sustainable and environment-
ing the mining of valuable ores (Smith et al., 2013). Most of the friendly society, two potential approaches for the remediation of
AMD treatment systems use limestone or lime as acid neutralizing gypsum waste are being investigated: material recycling and
agents, which results in the production of gypsum sludge (Kalin chemical recycling. Research with regards to material recycling
et al., 2006; Johnson and Hallberg, 2005). Long-term storage and has mostly been directed toward the utilization of gypsum waste
maintenance of gypsum waste present economic as well as poten- in the conditioning of alkaline soil (Gupta and Abrol, 1991) and
tial environmental concerns. The main environmental concerns the preparation of hardened gypsum products (Singh and Garg,
associated with waste gypsum dumps and disposal sites are the 2000).
In line with chemical recycling, the Council for Scientific and
Industrial Research (CSIR, Pretoria, SA) has developed a novel tech-
⇑ Corresponding authors at: DST/CSIR National Centre for Nanostructured Mate- nological process for the conversion of gypsum waste into high-
rials, Council for Scientific and Industrial Research, PO Box 395, Pretoria 0001, South value elemental sulphur, while a low-grade CaCO3 is generated
Africa (M. de Beer), Council for Geoscience, Private Bag X112, Pretoria 0001, South
as a by-product (De Beer et al., 2014; Mbhele et al., 2009;
Africa (F.J. Doucet).
E-mail addresses: mdebeer@csir.co.za (M. de Beer), fdoucet@geoscience.org.za Nengovhela et al., 2007; Maree, 2005). Sulphur is a key raw mate-
(F.J. Doucet). rial for many manufacturing industries such as fertilizer, acids,

http://dx.doi.org/10.1016/j.wasman.2015.08.023
0956-053X/Ó 2015 Published by Elsevier Ltd.

Please cite this article in press as: de Beer, M., et al. Synthesis of high-purity precipitated calcium carbonate during the process of recovery of elemental
sulphur from gypsum waste. Waste Management (2015), http://dx.doi.org/10.1016/j.wasman.2015.08.023
2 M. de Beer et al. / Waste Management xxx (2015) xxx–xxx

steel, petroleum, insecticides, and explosives (Jin et al., 2010). The 2008; Lazzeri et al., 2005; Li et al., 2002). The use of high-purity
CSIR technology involves three steps, namely, the thermal reduc- PCC can also benefit flue gas desulphurization technologies.
tion of calcium sulphate to calcium sulphide (Eq. (1)) at 1050 °C, Valuable, high-purity gypsum will precipitate from the reaction
the production of hydrogen sulphide from the calcium sulphide of pure CaCO3 with SO2, which is required for the production of
(Eq. (2)) at ambient conditions, and subsequently, the conversion gypsum boards. Because of the higher market price of pure
of hydrogen sulphide to elemental sulphur (Eq. (3)) via the com- CaCO3, the gypsum waste treatment process with the production
mercially available chemical catalytic Clauss process (Mark et al., of high-purity CaCO3 could be more economically feasible than
1978). the simple one-step or direct aqueous CaS carbonation step.
Thermal reduction of solid gypsum to produce a solid calcine Simultaneously, major benefits in terms of waste reduction are
product: realised.
This study focused on the synthesis and characterisation of
CaSO4  2H2 O ðsÞ þ 2C ðsÞ ! CaS ðsÞ þ 2CO2 ðgÞ þ 2H2 O ðlÞ ð1Þ high-purity CaCO3 from gypsum waste in a laboratory-scale exper-
Carbonation of the calcine product to produce hydrogen sul- imental set-up.
phide and calcium carbonate:
2. Experimental section
CaS ðsÞ þ H2 O ðlÞ þ CO2 ðgÞ ! H2 S ðgÞ þ CaCO3 ðsÞ ð2Þ
Recovery of elemental sulphur from the hydrogen sulphide gas: 2.1. Feedstock

2H2 S ðgÞ þ O2 ðgÞ ! 2S ðsÞ þ 2H2 O ðlÞ ð3Þ A calcium sulphide-rich calcine powder (hereafter denominated
Although the overall process successfully produced high-purity as CaSr) was produced from gypsum waste generated at an acid
elemental sulphur (Nengovhela et al., 2007), the study by de Beer mine water neutralisation plant in South Africa. Thermal reduction
et al. (2014) showed that the direct carbonation of calcium sul- of gypsum waste to CaSr using duff coal as the reducing agent was
phide (Eq. (2)) only yielded low-purity CaCO3 products (i.e. performed in a pilot-scale electrically-heated kiln using an existing
86 mass% as CaCO3). The insight gained on the solution chemistry process (Ruto et al., 2011). At completion of the reduction process,
and the characteristics of the carbonate products from the direct CaSr was stored in a nitrogen environment to avoid spontaneous
aqueous CaS carbonation process served as a basis for the develop- oxidation to CaSO4 (García-Calzada et al., 2000). The procedures
ment of a two-step or indirect carbonation process route for the used for the thermal reduction of gypsum waste and the character-
production of high-purity CaCO3 (i.e. >99 mass% as CaCO3). An out- isation (i.e. XRD and XRF) of gypsum waste and formed CaSr were
line of the indirect CaS carbonation process step is schematically described elsewhere (de Beer et al., 2014).
illustrated in Fig. 1 (the highlighted area represents the focus of Gaseous hydrogen sulphide (100% H2S; Pure Gas, South Africa)
this study). was used in all experiments to induce the dissolution of CaSr.
The feasibility of the gypsum technology can be enhanced if car- Gaseous carbon dioxide (100% CO2; Air Liquide, South Africa) was
bonate products of higher value can be recovered. In general, used to strip H2S gas from solution and to stimulate the carbona-
CaCO3 is a versatile material with a variety of uses. However, the tion of solubilized calcium derived from CaSr dissolution to form
application and uses of the low-purity, low-value CaCO3 is limited calcium carbonate.
to acid mine water neutralisation, cement manufacturing and met-
allurgical applications such as fluxing agent in steel making. High- 2.2. Experimental procedures
purity CaCO3 or precipitated calcium carbonate (PCC) has long
been recognized as a versatile additive and its importance is well All experiments were performed in a 3-L (for CaSr dissolution)
known due to its large range of industrial applications: e.g. as poly- and 1-L (for carbonation of Ca-rich leachate solution) Perspex stir-
mer composite; filler for rubber and ceramics; additives for plas- red tank, batch reactors. The two reactors were equipped with four
tics, paper, ink, cosmetics, toothpastes, detergents, a component equally spaced baffles and a sparger with small diameter (<1 mm)
of pharmaceuticals and foodstuffs as well as numerous uses in openings for the introduction of H2S or CO2 gas. A mechanical over-
electronics and catalysis (Donate-Robles and Martín-Martínez, head stirrer (RW 20 digital from IKAÒ-Werke GmbH & Co. KG,
2011; Shen et al., 2010, 2008; Gao et al., 2009; Gorna et al., Germany) and a Rushton turbine impeller (manufactured by

Sulphur recovery

2H2S + O2 → 2S + 2H2O Elemental Sulphur

Waste
Coal Aqueous stream
gypsum
H 2S
H2S stripping /
Thermal reducon CaS dissoluon Separaon Calcium carbonaon Separaon
CaSO4.2H2O + 2C → CaS CaS solids CaS + H2O + H2S → H2O Ca(HS)2 + CO2 + H2O →
+ 2CO2 + H2O + Ca(HS)2 CaCO3 ↓ + 2H2S ↑

CO2

Residue High-grade CaCO3

Fig. 1. Process flow diagram for the two-step, indirect CaS carbonation process using H2S gas for CaS dissolution and CO2 gas for CaCO3 precipitation.

Please cite this article in press as: de Beer, M., et al. Synthesis of high-purity precipitated calcium carbonate during the process of recovery of elemental
sulphur from gypsum waste. Waste Management (2015), http://dx.doi.org/10.1016/j.wasman.2015.08.023
M. de Beer et al. / Waste Management xxx (2015) xxx–xxx 3

Manten Engineering, South Africa) were used for mixing. The Universal Attenuated Total Reflectance (ATR) sampling accessory
Rushton turbine impeller had six vertical blades fixed onto a disk. and a diamond crystal. The spectral study was extended over the
A rotameter (Fisher & Porter, different models for H2S and CO2 gas) range 400–4000 cm1, with a resolution of 4 cm1 and 32 accumu-
was used to control the gas flow-rate. All experiments were con- lations. The spectra were analysed using the Spectrum 100 soft-
ducted at atmospheric pressure and room temperature. ware (Perkin Elmer).
Calcine (CaSr) was dispersed in distilled water (CaSr/water ratio The specific surface area of the products was measured by the
of 5%) to obtain a pre-determined slurry concentration. After Brunauer–Emmett–Teller (BET) method of N2 adsorption–desorp-
30 min of continuous mixing at 300 min1, H2S gas was introduced tion at 196 °C with an ASAP 2010 surface area analyzer
at a constant flow-rate into the slurry to induce CaSr dissolution. (Micrometrics BET, USA). Prior to analysis, the samples were pre-
The pH, electrical conductivity and temperature of the suspension conditioned at 120 °C in vacuum for 4 h to remove the humidity
were logged at 5 s intervals using a Hanna HI 2829 multi- and contaminants adsorbed to the particle surfaces.
parameter logger (Hanna Instruments, South Africa) to collect A helium gas displacement method with a pycnometer
time-series data on the dynamics of the system. The reaction was (AccuPyc II 1340, Micromeritics, USA) was used to measure the
terminated by shutting off the H2S gas stream when the pH and true density (qs) of the carbonate products.
electrical conductivity remained unchanged for 15–20 min. Upon
completion, the suspension was removed from the reactor by vac- 3. Results and discussion
uum filtration using Whatman grade no. 1 filter paper. The filter
cake was thoroughly rinsed with water and dried at mild temper- 3.1. Characterisation of gypsum waste and calcine material
ature (i.e. 60 °C; to prevent thermal conversion of gypsum
(Ballirano and Melis, 2009)) for 24 h. The mineralogical composi- The untreated gypsum waste used in this study was charac-
tion of the undissolved residues was analysed using XRD and their terised and described elsewhere (de Beer et al., 2014). In brief, it
particle size distribution was determined using the laser diffrac- featured a mean diameter D (m, 0.5) of 43.9 lm with 10% of the par-
tion method (Horiba LA-950). The filtrate (Ca2+-rich Ca(HS)2 solu- ticles smaller than 4.0 lm and 90% of the particles smaller than
tion) was transferred into the 1-L Perspex reactor for further 176.7 lm. Its bulk mineralogical composition consisted of gypsum
reaction with CO2. (CaSO42H2O; 96.9%), which co-existed with magnesite (MgCO3;
After 10 min of continuous mixing of the filtrates (Ca(HS)2 solu- 3.1%). In contrast to our previous study (de Beer et al., 2014) where
tions containing 450 mmol/L, 900 mmol/L and 1800 mmol/L Ca2+ gypsum waste had only been partly converted to a calcium
respectively) from the CaS dissolution step in the 1-L reactor, sulphide-rich calcine powder (CaSr) containing 50.7% of the min-
CO2 gas was introduced at a constant flow-rate via a sparger in eral oldhamite (i.e. CaS), this study benefited from an optimised
order to precipitate high-grade CaCO3. The pH, electrical conduc- thermal reduction process. The latter generated a CaSr sample of
tivity and temperature of the suspension in the reactor were higher purity (83.0% oldhamite). Less abundant mineral phases
recorded as described earlier. Samples of the Ca-rich solutions sub- included hydroxyapatite (Ca5(PO4)3(OH); 8.2%), quartz (SiO2;
jected to carbonation were also collected from the reactor at regu- 4.3%), anhydrite (CaSO4; 4.2%), and lime (CaO; 0.3%). The level of
lar intervals and were filtered using 0.45 lm PALL acrodisc PSF purity of the CaSr sample was further confirmed by a wet analytical
GxF/GHP membranes (Microsep (Pty) Ltd, South Africa). The fil- method (Procedure 4500-S2/Iodometric method described in
trates (undiluted, acidified samples) were analysed for their cal- Standard Methods; Clesceri et al., 1989), which reported a CaS con-
cium contents by inductively coupled plasma mass spectrometry tent of 81.7 ± 0.8 mass%.
(ICP-MS) at an accredited laboratory (Consulting and Analytical
Services, CSIR, Pretoria, South Africa). The reaction was terminated 3.2. Aqueous dissolution of CaSr
when the pH and electrical conductivity remained unchanged for
10–15 min, indicating the completion of the reaction. As conveyed in our previous paper (de Beer et al., 2014), CaS is a
Immediately upon completion of the experimental run, the final sparingly soluble salt in water, as indicated by its low solubility
suspension was removed from the reactor by vacuum filtration ranging from 0.125 g/L to 1.0 g/L (Dean, 1992; Perry and Green,
using 0.45 lm Millipore HA membranes (Microsep (Pty) Ltd, 1984; Weast, 1972). However, the extent of its dissolution can be
South Africa) and washed twice with distilled water. The filter substantially improved by sparging H2S gas to aqueous CaS sus-
cakes were dried at 60 °C for 24 h. pensions. In order to confirm this effect, CaSr was initially sus-
pended in water (without the addition of H2S gas) and
2.3. Characterisation of solid materials continuously mixed for 30 min. The solution pH rapidly increased
and stabilized at values above 10 within the first 2 min of stirring.
The mineralogical composition of the gypsum waste, calcine This was accompanied by an increase in solution conductivity
(CaSr) and the carbonate products were analysed by XRD (1.7 mS/cm). This confirmed our previous observations made with
(PANalytical X’Pert Pro powder diffractometer). The phases were CaSr of lower purity (i.e. 50.7% CaS; de Beer et al., 2014). In addi-
identified using the X’Pert Highscore plus software. Relative phase tion, the high sulphur content (395 mmol/L as S) of CaSr dispersed
amounts (wt%) were estimated using the Rietveld method. in water and the low amount (4 mmol/L as S) of dissolved sulphur
High-vacuum scanning electron micrographs were collected measured in solution after the 30 min of mixing in the absence of
using a JEOL JSM7500 microscope to obtain information on mor- H2S gas confirmed the low solubility of CaS in water.
phologies and size distribution of all the solid materials. Samples H2S gas was then bubbled through CaSr suspensions in order to
were dispersed on carbon tape and sputter-coated with a thin, con- promote the aqueous dissolution of CaS particles. It was antici-
ductive layer of gold using a Emitech K950X sputter coater. pated that H2S gas would first dissolve in water to produce aque-
Thermogravimetric analysis of the dried carbonate powders ous H2S, which subsequently would react with CaS to form
was performed using the TGA Q500 (TA Instruments) in a nitrogen water-soluble Ca(HS)2 (248.7 g/L; Seidell and Linke, 1940). In the
atmosphere. Samples were heated from room temperature to 900 ° presence of H2S, CaS dissolution would therefore proceed accord-
C at a heating rate of 10 °C/min. ing to Eq. (4).
Infrared spectra were recorded at room temperature with a
CaS ðsÞ þ H2 O ðlÞ þ H2 S ðaqÞ ! CaðHSÞ2 ðaqÞ þ H2 O ðlÞ
Perkin–Elmer Spectrum 100 Fourier Transform Infrared
Spectrometer (FTIR) equipped with a Perkin–Elmer Precisely
DH25 C ¼ 44:8 kJ ð4Þ

Please cite this article in press as: de Beer, M., et al. Synthesis of high-purity precipitated calcium carbonate during the process of recovery of elemental
sulphur from gypsum waste. Waste Management (2015), http://dx.doi.org/10.1016/j.wasman.2015.08.023
4 M. de Beer et al. / Waste Management xxx (2015) xxx–xxx

Following the addition of H2S gas (time = 0 min) to the aqueous the total amount of calcium added as CaS (575 mmol/L or 69.1 g
CaS suspension, immediate temporal changes (i.e. increase/ as Ca), 91.7% dissolved upon H2S addition to form a Ca(HS)2 solu-
decrease) in solution pH, electrical conductivity, dissolved calcium tion containing 530 mmol/L (or 63.7 g) as Ca 45 mmol/L (or 5.4 g
concentration and temperature were observed. The conductivity, as Ca) of the added Ca was lost to the residue.
temperature and dissolved calcium exhibited similarly-shaped The dissolution of CaS using H2S gas generated a calcium-rich
profiles, with all three parameters having reached their maxima solution (to be subsequently carbonated), and a dark-grey solid
after about 20 min of reaction (Fig. 2). The solution conductivity residue (waste). Approximately 0.38 kg of this residue was gener-
(Fig. 2a), temperature (Fig. 2b) and dissolved calcium (Fig. 2c) ated for every 1.0 kg of CaSr, or for every 2 kg of waste gypsum pro-
increased over the first 20 min before levelling off and stabilizing cessed. This represented a reduction of ca. 80% in mass of waste to
at about 54 mS/cm, 28 °C and 530 mmol/L respectively. The forma- be discarded, compared to the starting gypsum waste mass. The
tion of the plateau was much more rapid (ca. 20 min; Fig. 2) than residue was characterised by a mean diameter D (m, 0.5) of
that observed previously for direct carbonation in the presence of 14.3 lm with 10% of the particles smaller than 6.34 lm and 90%
CO2 and the absence of H2S (200–210 min; de Beer et al., 2014). of the particles smaller than 64.39 lm. The primary mineral phases
The change in temperature from 21.7 to 28.3 °C (DT = 6.6 °C) con- identified in the residue were anhydrite (CaSO4; 35.7%), apatite
firmed the exothermic nature of CaS dissolution. The increase in (Ca5(PO4)3OH; 27.9%), quartz (SiO2; 13.6%) and two carbonate
conductivity could be ascribed to the absorption of H2S gas and polymorphs, calcite and vaterite (CaCO3; 12.3%). Less abundant
the subsequent formation of aqueous H2S, as well as to the disso- mineral phases included fluorite (CaF2; 5.7%) and bassanite
lution of CaS present in CaSr. The gradual increase in conductivity (CaSO4½H2O; 4.4%). Apart from the carbonate phases, the other
values was therefore a first indirect indication of the formation of primary minerals were present in the starting CaSr powder and
Ca(HS)2 in solution. Over the pH range 7.6–11.8, HS is the domi- most probably did not react during the H2S-facilitated dissolution
nant sulphide species (Suleimenov and Krupp, 1994), which step.
implies that during the dissolution stage, little, if any S is stripped
from the reactor system. During the same period, the solution pH
3.3. Carbonation of calcium-rich solutions
(Fig. 2d) initially dropped sharply from 11.9 to 9.6 within the first
minute, after which it continued decreasing more gradually down
Experiments of aqueous dissolution of CaSr were repeated using
to ca. 7.6 at 20 min, levelled off and remained constant from 20 to
distinct CaSr/water ratios of 4%, 8% and 16% (or 97.2 g, 194.4 g and
40 min of reaction time. The assumption was made that complete
388.8 g CaSr in 3 L of water) in order to produce Ca(HS)2 solutions
dissolution of CaS particles would be indicated by the levelling-off
containing 450 mmol/L, 900 mmol/L and 1800 mmol/L Ca2+
of all the salient parameters; this assumption was supported by the
respectively. At completion of these dissolution experiments, the
levelling off of dissolved calcium (Fig. 2c) and the observed reduc-
aqueous Ca2+-rich filtrates were subjected to a carbonation reac-
tion in solution temperature after 20 min of reaction (Fig. 2b). Of
tion using flow rates of 1.05 (for 4% and 8% CaSr/water ratios)

60 30
(a) (b)
50
Conductivity (mS/cm)

28
Temperature (°C)

40
26
30
24
20

22
10

0 20
0 10 20 30 40 0 10 20 30 40
Time (min) Time (min)

600 100
(c) 12 (d)
Dissolved calcium (mmol/L)

500 80
Dissolved calcium (%)

% dissolved calcium 11
400
60
10
300
pH

40
9
200

20 8
100

0 0 7
0 10 20 30 40 0 10 20 30 40
Time (min) Time (min)

Fig. 2. Solution conductivity (a), temperature (b), dissolved calcium (mmol/L on Y1 and % on Y2 axis) (c) and pH (d) profiles of the calcine slurry in equilibrium with distilled
water upon H2S addition (CaS slurry containing 575 mmol/L as S; initial pH: 11.9; gas flow: 8.4 L/min/kg calcine; stirring rate: 700 min1; a time of 0 min on the x-axis
corresponds to the start of the addition of H2S to solution).

Please cite this article in press as: de Beer, M., et al. Synthesis of high-purity precipitated calcium carbonate during the process of recovery of elemental
sulphur from gypsum waste. Waste Management (2015), http://dx.doi.org/10.1016/j.wasman.2015.08.023
M. de Beer et al. / Waste Management xxx (2015) xxx–xxx 5

and 1.75 (for 16% CaSr/water ratio) L/min/mol Ca2+ to ensure that are the in-plane (t4 mode) and the out-of plane (t2 mode) vibra-
CO2 was present in excess in relation to the concentration of dis- tional bands at 713 cm1 and 877 cm1, respectively, and the
solved Ca2+. anti-symmetry stretch (t3 mode) at around 1400 cm1 and sym-
Upon introduction of CO2 gas, CaCO3 was expected to precipi- metric stretching (t1 mode) at around 1800 cm1 (Xu and
tate from solution according to Eq. (5): Poduska, 2014; Menahem and Mastai, 2008). The IR spectra con-
firmed the presence of the single calcite phase by the identification
CaðHSÞ2 ðaqÞ þ CO2 ðgÞ þ H2 O ðlÞ of the characteristic t4 band of calcite at 713 cm1 and the absence
! CaCO3 ðsÞ þ 2H2 S ðgÞ DH25 C ¼ 7:0 kJ ð5Þ of the characteristic t4 bands of vaterite at 745 cm1 and aragonite
at 700 cm1.

3.3.1. Analysis of CaCO3 polymorphism 3.3.2. Purity analysis


The phase identification of the carbonate products was carried The thermogravimetric analysis curves (Fig. 5) illustrate the
out by XRD. XRD patterns of the samples produced at increasing weight loss of the carbonate products during thermal decomposi-
initial dissolved Ca2+ concentrations are shown in Fig. 3. tion over the temperature range 25–900 °C. All three products
Diffraction patterns of the products exhibited a significant peak were stable below 100 °C. The first step was observed at about
at 29.3° (2h), which is characteristic of crystalline calcite (Chen 100 °C; it was characterised by a small weight loss (<0.6%), which
et al., 2012). In comparison with the standard JCPDS file (calcite: was attributed to loss of moisture from the samples. A second step
83-0586), the diffraction peaks were indexed as rhombohedral cal- was identified in the temperature range 600 to 730–760 °C; this
cite. Based on the Scherrer equation, the crystallite size is reversely weight loss was rapid, pronounced, and corresponded to the
proportional to the width of the peak at half maximum of the decomposition of carbonate products with the accompanying
diffraction line on the 2h scale (Langford and Wilson, 1978). release of CO2 (Wang et al., 1999). This weight loss equated to
According to the equation on the 104 peak located at 2h = 29.3°, approximately 43.6% of the total mass. The sample weights
the average grain size of the calcite crystallite of all the products remained unchanged upon heating above 760 °C. Calcium carbon-
were identical at 834 Å or 83.47 nm. The sharp and strong peaks ate thermally decomposes according to the following reaction:
further confirmed that the products were well-crystallized
(Fig. 3a–c). The bulk mineral composition, as determined by CaCO3 ðsÞ ! CaO ðsÞ þ CO2 ðgÞ ðloss of carbon dioxideÞ
Rietveld refinement of XRD diffraction data, was primarily calcite 1
DH25 C ¼ 182:1 kJ mol ð6Þ
(99.5%) with traces of quartz (0.5%). According to XRD, calcite
was therefore the only CaCO3 polymorph obtained at the end of Additional TGA experiments were performed on these three
each reaction; no vaterite or aragonite was detected under these samples to obtain additional quantification between the weight
experimental conditions. fractions. For this purpose, during heating, the samples were kept
The formation of a single polymorph of CaCO3 was further con- isothermally at 105 °C, 400 °C and 850 °C for 15 min. The weight
firmed by FT-IR analysis (Fig. 4). Infrared spectroscopy detects the fraction (Dm400850°C) based on dry weight (m105°C) was used to
vibration characteristics of chemical functional groups in samples. calculate the calcium carbonate content, which was expressed as
When an infrared light interacts with the sample, chemical bonds wt% of CO2, using equation Eq. (7).
will stretch, contract and bend. As a result, a chemical functional
CO2 ðwt%Þ ¼ Dm400—850 C =m105 C  100 ð7Þ
group tends to adsorb infrared radiation in a specific wavenumber
range regardless of the structure of the rest of the molecule. FT-IR Analysis of the thermal decomposition of the carbonate prod-
is therefore a useful tool for the identification of calcium carbonate ucts synthesised at initial Ca2+ concentrations of 450 mmol/L,
polymorphs due to the differences in their carbonate ions, CO2 3 900 mmol/L and 1800 mmol/L resulted in actual mass loss of
(Islam et al., 2011). The spectral data obtained for the samples 43.5%, 43.7% and 43.6%, respectively, which is very similar to the
revealed the characteristic transmittance peaks of calcite, which theoretical mass loss expected for 100% pure CaCO3 (i.e. 43.9%;

8000 (a)
6000
4000
2000
0
10 20 30 40 50 60 70
Intensity (counts)

8000 (b)
6000
4000
2000
0
10 20 30 40 50 60 70
8000 (c)
6000
4000
2000
0
10 20 30 40 50 60 70
2 Theta (°)

Fig. 3. XRD diffraction patterns of high-grade CaCO3 produced at initial Ca2+ concentrations of (a) 450 mmol/L, (b) 900 mmol/L and (c) 1800 mmol/L.

Please cite this article in press as: de Beer, M., et al. Synthesis of high-purity precipitated calcium carbonate during the process of recovery of elemental
sulphur from gypsum waste. Waste Management (2015), http://dx.doi.org/10.1016/j.wasman.2015.08.023
6 M. de Beer et al. / Waste Management xxx (2015) xxx–xxx

100
90 (a)
80
70
60
2000 1800 1600 1400 1200 1000 800 600
Transmittance (%)
100
90 (b)
80
70
60
2000 1800 1600 1400 1200 1000 800 600
100
90 (c)
80
70
60
2000 1800 1600 1400 1200 1000 800 600
-1
Wavenumber (cm )

Fig. 4. FT-IR transmission spectra of high-grade CaCO3 produced at initial Ca2+ concentrations of (a) 450 mmol/L, (b) 900 mmol/L and (c) 1800 mmol/L.

100 Increasing the initial Ca2+ concentrations from 450 to


(a) Adsorbed water: 0.39% 1800 mmol/L Ca2+ in the filtrates had therefore little effect on the
90 mineral composition, purity and percentage yield of the formed
Mass loss:
80 43.54% carbonate products (Table 1). The nearly identical CaCO3 yield
obtained confirmed that CO2 had been present in excess in relation
70
to dissolved Ca2+ concentration.
60 759 °C The product generated from a 1800 mmol/L (as Ca) solution
50 decomposed slightly more rapid (Fig. 5c; 600–730 °C) than the
100 200 300 400 500 600 700 800 900 products obtained from solutions containing 900 or 450 mmol/L
Weight (% of initial mass)

100 (as Ca) (Fig. 5a and b; 600–760 °C). The slight difference could
(b) Adsorbed water: 0.54% not be explained by the polymorph composition, crystallite size
90
Mass loss:
and purity of the products. It was also required to characterise
80 43.71% the products in terms of their physical properties.
70

60 756 °C 3.3.3. Structural and morphological analysis


The effect of the initial dissolved Ca2+ concentration in solution
50
100 200 300 400 500 600 700 800 900 on the surface characteristics and morphology of CaCO3 crystals
was also determined. SEM observations indicated precipitates of
100
micron-sized, interpenetrated rhombohedral cubes of calcite
90
(c) Adsorbed water: 0.57%
(Fig. 6), with a notable decrease in particle size with increasing ini-
80 Mass loss: tial Ca2+ concentration. A laser diffraction technique was used to
43.62% compare the particle size characteristics of the precipitated solids.
70
Fig. 7a shows the cumulative volume fractions and Fig. 7b the par-
60 732 °C ticle size frequency distributions. Data in Table 2 show the mea-
50 sured true density, surface area and particle sizes of the
100 200 300 400 500 600 700 800 900 products. With increasing initial Ca2+ concentration, the specific
Temperature (°C) surface area of the precipitates increased as the mean particle size
diameter became smaller. The particles of smallest size
Fig. 5. TGA curves for high-grade CaCO3 produced at initial Ca2+ concentrations of
(a) 450 mmol/L, (b) 900 mmol/L and (c) 1800 mmol/L.
Table 1
Quantitative XRD analysis and yields of high-grade CaCO3 produced at different initial
Frost et al., 2009). The purity of CaCO3 in the products was calcu- Ca2+-concentrations during the carbonation step.

lated using Eq. (8), where MW denotes molecular weights and Initial Ca2+ concentration mmol/L 450 900 1800
CO2 (wt%) represents the weight loss of the sample. Mineral composition
Calcite mass% 99.5 99.5 99.4
PCaCO3 ð%Þ ¼ CO2 ðwt%Þ  MWCaCO3 =MWCO2 ð8Þ Quartz mass% 0.5 0.5 0.6
The purity of the CaCO3 in the formed products was consistently CaCO3 yield
higher than 99.0% and confirmed the high purity of the synthesized Actual mass g/100 g CaSr 72.2 ± 0.7 71.8 ± 1.0 70.2 ± 0.2
Yield % 91.6 ± 0.8 90.8 ± 1.2 88.8 ± 0.2
carbonate products suggested by XRD (Table 1).

Please cite this article in press as: de Beer, M., et al. Synthesis of high-purity precipitated calcium carbonate during the process of recovery of elemental
sulphur from gypsum waste. Waste Management (2015), http://dx.doi.org/10.1016/j.wasman.2015.08.023
M. de Beer et al. / Waste Management xxx (2015) xxx–xxx 7

Fig. 6. SEM images (top row 1000 and bottom row 5000 magnification) of high-grade CaCO3 produced at initial Ca2+ concentrations of (a) 450 mmol/L, (b) 900 mmol/L and
(c) 1800 mmol/L.

(a) These findings were in agreement with the theoretical models of


Cumulative volume

100
precipitation by which an increase in initial Ca2+ concentration
fraction (%)

80 (i) 450 mmol/L gives rise to nucleation of a larger number of smaller crystals
(ii) 900 mmol/L
60 (iii) 1800 mmol/L (Mumtaz and Hounslow, 2000).
40
20
iii ii i 3.3.4. Solution chemistry
Fig. 8 illustrates temporal changes in solution pH and dissolved
0
0.1 1 10 100 1000 Ca2+ concentration during the carbonation of the filtrates charac-
terised by varying initial dissolved Ca2+ concentrations. It must
Diameter (µm)
be noted that the pH at the end of the dissolution stage was around
(% particles by volume)

25 7.5 while H2S gas was introduced. As soon as the gas was shut off
(b) and the solid precipitate removed, the pH increased rapidly to
20 ii
about 10.5, presumably due to the volatilisation of excess H2S.
Frequency

15 (i) 450 mmol/L When adding CO2 (time = 0 min), the pH (Fig. 8, bottom row)
(ii) 900 mmol/L
10 (iii) 1800 mmol/L dropped sharply from >10.5 to approximately 8.5 within a very
iii
i short period of time (<2 min), after which it continued decreasing
5
more gradually before stabilizing at approximately 6.5 at the end
0 of the CaCO3 precipitation reaction.
0.1 1 10 100 1000
Plateauing of the dissolved calcium concentration indicated the
Diameter (µm)
end of the CaCO3 formation. The descending parts of dissolved Ca
Fig. 7. Particle size distribution curves of cumulative volume fraction (a) and concentration vs. time curves correspond to the consumption of
frequency (b) for high-grade CaCO3 produced at various initial Ca2+ concentrations. dissolved Ca and the concurrent formation of CaCO3. It was found
that all three reactions had been completed within a similar
(Figs. 6c, 7aiii, biii and 8) decomposed faster (Fig. 5c) than larger amount of time (50–55 min). For this reason and because the initial
particles (Figs. 6a and b, 7ai and bi, 8 and 5a and b), most probably dissolved Ca concentrations had been dissimilar, the slope of the
because smaller-sized particles have a greater surface area per unit lines did not represent the rate of CaCO3 precipitation, but was
mass than larger particles (Table 2). The products generated at dif- rather representative of the differences in the amount of Ca being
ferent initial dissolved Ca2+ concentrations exhibited densities of converted to carbonates. Fig. 8 depicts calculated values obtained
2.7 g/cm3, which is characteristic of calcite (Plummer and for the slopes of the calcium vs time profiles. The higher values
Busenberg, 1982). obtained for starting filtrates containing 1800 mmol/L Ca com-
The degree of aggregation of the precipitated particles also cor- pared to those containing less Ca indicated that much more Ca
related with the increase in initial Ca2+ concentration, as shown had been consumed during the carbonation stage of the former
from the SEM images of the corresponding samples (Fig. 6). solutions.

Table 2
Physical properties of high-grade CaCO3 produced at increasing levels of Ca2+ concentrations in the starting filtrates.

Initial [Ca2+] True density (g/cm3) BET surface area (m2/g) Particle size distribution
D (v, 0.1) (lm) D (v, 0.5) (lm) D (v, 0.9) (lm)
(a) 450 mmol/L 2.710 ± 0.003 0.30 ± 0.01 27.73 44.32 82.30
(b) 900 mmol/L 2.710 ± 0.005 0.40 ± 0.02 23.31 32.36 46.15
(c) 1800 mmol/L 2.720 ± 0.007 0.79 ± 0.03 13.16 24.65 42.10

Please cite this article in press as: de Beer, M., et al. Synthesis of high-purity precipitated calcium carbonate during the process of recovery of elemental
sulphur from gypsum waste. Waste Management (2015), http://dx.doi.org/10.1016/j.wasman.2015.08.023
8 M. de Beer et al. / Waste Management xxx (2015) xxx–xxx

2000 (a) 2000 (b) 2000 (c)

-36.2 ± 0.1 mmol/L/min

Calcium (mmol/L)
1500 1500 1500 2
(R = 0.9977)

1000 1000 -15.7 ± 0.1 mmol/L/min 1000


2
(R = 0.9988)
-9.5 ± 0.1 mmol/L/min
2
(R = 0.9993)
500 500 500

0 0 0
0 20 40 60 80 0 20 40 60 80 0 20 40 60 80
Time (min) Time (min) Time (min)

11 11 11
(d) (e) (f)
10 10 10

9 9 9
pH

8 8 8

7 7 7

6 6 6

0 20 40 60 80 0 20 40 60 80 0 20 40 60 80
Time (min) Time (min) Time (min)

Fig. 8. Dissolved calcium (a) and pH (d) profiles of the carbonation reaction at the initial Ca2+ of 450 mmol/L, initial pH 10.8 and gas flow rate of 1.05 L/min/mol Ca2+.
Dissolved calcium (b) and pH (e) profiles of the carbonation reaction at the initial Ca2+ of 900 mmol/L, initial pH 11.5 and gas flow rate of 1.05 L/min/mol Ca2+. Dissolved
calcium (c) and pH (f) profiles of the carbonation reaction at the initial Ca2+ of 1800 mmol/L, initial pH 10.5 and gas flow rate of 1.75 L/min/mol Ca2+.

4. Conclusions References

The potential for the synthesis of high-purity CaCO3 from indus- Ballirano, P., Melis, E., 2009. Thermal behaviour and kinetics of dehydration of
gypsum in air from in situ real-time laboratory parallel-beam X-ray powder
trial gypsum waste was demonstrated in this study. High-grade diffraction. Phys. Chem. Miner. 36 (7), 391–402.
CaCO3 (i.e. >99.5 mass% as CaCO3) or PCC was obtained during Chen, Z., Yang, B., Nan, Z., 2012. Modification of calcium carbonate crystals growth
the indirect aqueous carbonation of CaS, an intermediate product by dibenzoic acids. Mater. Chem. Phys. 132 (2–3), 601–609.
Clesceri, L.S., Greenberg, A.E., Trussel, R.R. (Eds.), 1989. Standard Methods for the
in the process of the recovery of elemental sulphur from gypsum Examination of Water and Wastewater, 17th ed. American Public Health
waste. We are now optimising the process further to generate Association, Washington.
sub-micron, high-purity CaCO3 particles. The effect of initial Ca2+ De Beer, M., Maree, J.P., Liebenberg, L., Doucet, F.J., 2014. Conversion of calcium
sulphide to calcium carbonate during the process of recovery of elemental
concentration in solution, which is a function of the solid–liquid sulphur from gypsum waste. Waste Manage. 34, 2373–2381.
ratio (CaSr/water ratio during the dissolution step) on the formed Dean, J.A. (Ed.), 1992. Lange’s Handbook of Chemistry, 14th ed. McGraw-Hill, New
carbonate products, was investigated. The effect on the final car- York.
Donate-Robles, J., Martín-Martínez, J.M., 2011. Addition of precipitated calcium
bonate products in terms of mineral composition, purity, density
carbonate filler to thermoplastic polyurethane adhesives. Int. J. Adhes. Adhes.
and percentage yield were similar. The results on the mineralogical 31 (8), 795–804.
analyses also showed that calcite was the only CaCO3 polymorph Frost, R.L., Hales, M.C., Martens, W.N., 2009. Thermogravimetric analysis of selected
group (II) carbonate minerals – Implication for the geosequestration of
obtained. Rhombohedral structured calcite can find application in
greenhouse gases. J. Therm. Anal. Calorim. 95 (3), 999–1005.
for instance the paper and pulp industry where sheet bulk, porosity Gao, Y., Liu, L., Zhang, Z., 2009. Mechanical performance of nano-CaCO3 filled
or some other property is of paramount importance. This process polystyrene composites. Acta Mech. Solida Sin. 22 (6), 555–562.
offers the significant advantage of producing two commercial- García-Calzada, M., Marbán, G., Fuertes, A.B., 2000. Decomposition of CaS particles
at ambient conditions. Chem. Eng. Sci. 55 (9), 1661–1674.
grade products (elemental sulphur, and PCC) from gypsum waste. Gorna, K., Hund, M., Vučak, M., Gröhn, F., Wegner, G., 2008. Amorphous calcium
carbonate in form of spherical nanosized particles and its application as fillers
Acknowledgments for polymers. Mater. Sci. Eng. A 477 (1–2), 217–225.
Gupta, R.K., Abrol, I.P. (Eds.), 1991. Salt-affected soils: their Reclamation and
Management for Crop Production. Springer, New York.
The authors thank the following organisations for financial sup- Islam, K.N., Bakar, M.Z.B.A., Noordin, M.M., Hussein, M.Z.B., Rahman, N.S.B.A., Ali, M.
port and/or technical/scientific input: THRIP (Technology and E., 2011. Characterisation of calcium carbonate and its polymorphs from cockle
shells (Anadara granosa). Powder Technol. 213 (1–3), 188–191.
Human Resource for Industry Programme of the National
Jin, S., Shuichang, Z., Guangyou, Z., Zhi, Y., Bin, Z., Anguo, F., Debin, Y., 2010.
Research Foundation (NRF)), Tshwane University of Technology Geological reserves of sulfur in China’s sour gas fields and the strategy of sulfur
(TUT), Northwest University (NWU), the Council for Scientific and markets. Petrol. Explor. Dev. 37 (3), 369–377.
Industrial Research (CSIR) and the Council for Geoscience (CGS). Johnson, D.B., Hallberg, K.B., 2005. Acid mine drainage remediation options: a
review. Sci. Total Environ. 338 (1–2), 3–14.
The authors are grateful to the DST/CSIR National Centre for Kalin, M., Fyson, A., Wheeler, W.N., 2006. The chemistry of conventional and
Nanostructured Materials characterisation facility staff, specifically alternative treatment systems for the neutralization of acid mine drainage. Sci.
K Selatile and Z Ndaba for their help with analyses. Total Environ. 366 (2–3), 395–408.

Please cite this article in press as: de Beer, M., et al. Synthesis of high-purity precipitated calcium carbonate during the process of recovery of elemental
sulphur from gypsum waste. Waste Management (2015), http://dx.doi.org/10.1016/j.wasman.2015.08.023
M. de Beer et al. / Waste Management xxx (2015) xxx–xxx 9

Langford, J.I., Wilson, A.J.C., 1978. Scherrer after sixty years: a survey and some new Minerals Industry Conference, 15–17 February 2011, The Southern African
results in the determination of crystallite size. J. Appl. Crystallogr. 11 (2), 102– Institute of Mining and Metallurgy, The Southern African Institute of Mining
113. and Metallurgy.
Lazzeri, A., Zebarjad, S.M., Pracella, M., Cavalier, K., Rosa, R., 2005. Filler toughening Seidell, A., Linke, W.F. (Eds.), 1940. Solubilities of Inorganic and Metal Organic
of plastics. Part 1 – The effect of surface interactions on physico-mechanical Compounds; a Compilation of Quantitative Solubility Data from the Periodical
properties and rheological behaviour of ultrafine CaCO3/HDPE nanocomposites. Literature. D. Van Nostrand Company Inc., New York.
Polymer 46 (3), 827–844. Shen, J., Song, Z.Q., Qian, X.R., Song, C.J., 2008. Preparation of chitosan-coated
Li, L., Collis, A., Pelton, R., 2002. A new analysis of filler effects on paper strength. J. papermaking grade PCC filler and its application. Chung-kuo Tsao Chih/China
Pulp Pap. Sci. 28 (8), 267–273. Pulp Paper 27 (1), 21–25.
Maree, J.P., 2005. Conversion of Sulphur-Containing Waste Material into a Sulphur- Shen, J., Song, Z., Qian, X., Yang, F., 2010. Carboxymethyl cellulose/alum modified
Containing Product. Australia: AU 2005201759 patent http://ip.com/patfam/en/ precipitated calcium carbonate fillers: preparation and their use in
35455786. papermaking. Carbohydr. Polym. 81 (3), 545–553.
Mark, H.F., Othmer, D.F., Overberger, C.G., Seaborg, G.T. (Eds.), 1978. Kirk-Othmer: Singh, M., Garg, M., 2000. Making of anhydrite cement from waste gypsum. Cem.
Encyclopedia of Chemical Technology (22, 276), third ed. Wiley-Interscience, Concr. Res. 30 (4), 571–577.
New York. Smith, L.J.D., Moncur, M.C., Neuner, M., Gupton, M., Blowes, D.W., Smith, L., Sego, D.
Mbhele, N.R., Van der Merwe, W., Maree, J.P., Theron, D., 2009. Recovery of Sulphur C., 2013. The Diavik waste rock project: design, construction, and
from Waste Gypsum. In: Abstracts of the International Mine Water Conference, instrumentation of field-scale experimental waste-rock piles. Appl. Geochem.
19–23 October 2009, pp. 622–630. 36, 187–199.
Menahem, T., Mastai, Y., 2008. Controlled crystallization of calcium carbonate Suleimenov, O.M., Krupp, R.E., 1994. Solubility of hydrogen sulfide in pure water
superstructures in macroemulsions. J. Cryst. Growth 310 (15), 3552–3556. and in NaCl solutions, from 20 to 320 °C and at saturation pressures. Geochim.
Mumtaz, H.S., Hounslow, M.J., 2000. Aggregation during precipitation from Cosmochim. Acta 58 (11), 2433–2444.
solution: an experimental investigation using Poiseuille flow. Chem. Eng. Sci. Tayibi, H., Choura, M., López, F.A., Alguacil, F.J., López-Delgado, A., 2009.
55 (23), 5671–5681. Environmental impact and management of phospho-gypsum. J. Environ.
Nengovhela, N.R., Strydom, C.A., Maree, J.P., Oosthuizen, S., Theron, D.J., 2007. Manage. 90 (8), 2377–2386.
Recovery of sulphur and calcium carbonate from waste gypsum. Water SA 33 Wang, L., Sondi, I., Matijevic, E., 1999. Preparation of uniform needle-like aragonite
(5), 741–747. particles by homogeneous precipitation. J. Colloid Interface Sci. 218 (2), 545–
Perry, R.H., Green, D. (Eds.), 1984. Perry’s Chemical Engineers’ Handbook, sixth ed. 553.
McGraw-Hill Book Company, New York. Weast, R.C. (Ed.), 1972. CRC Handbook of Chemistry and Physics, 53rd ed. The
Plummer, L.N., Busenberg, E., 1982. The solubilities of calcite, aragonite and vaterite Chemical Rubber Co., Cleveland, Ohio.
in CO2–H2O solutions between 0 and 90 °C, and an evaluation of the aqueous Xu, B., Poduska, K.M., 2014. Linking crystal structure with temperature-sensitive
model for the system CO3-CO2-H2O. Geochim. Cosmochim. Acta 46 (6), 1011– vibrational modes in calcium carbonate minerals. Phys. Chem. Chem. Phys. 16
1040. (33), 17634–17639.
Ruto, S., Maree, J.P., Zvinowanda, C.M., Louw, W.J., Kolesnikov, A.V., 2011. Thermal
studies on gypsum in a pilot-scale, rotary kiln. In: Water in the South African

Please cite this article in press as: de Beer, M., et al. Synthesis of high-purity precipitated calcium carbonate during the process of recovery of elemental
sulphur from gypsum waste. Waste Management (2015), http://dx.doi.org/10.1016/j.wasman.2015.08.023

You might also like