Q1+2019+Effects of Fly Ash Properties On Carbonation Efficiency in CO2 Mineralisation

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Fuel Processing Technology 188 (2019) 79–88

Contents lists available at ScienceDirect

Fuel Processing Technology


journal homepage: www.elsevier.com/locate/fuproc

Effects of fly ash properties on carbonation efficiency in CO2 mineralisation T


a,b,c b,⁎ c d d b
Long Ji , Hai Yu , Ruijie Zhang , David French , Mihaela Grigore , Bing Yu ,
Xiaolong Wange, Jianglong Yuf, Shuaifei Zhaoa,g,
⁎⁎

a
Department of Environmental Sciences, Macquarie University, North Ryde, NSW 2109, Australia
b
CSIRO Energy, Newcastle, NSW 2304, Australia
c
China University of Mining & Technology Beijing, Beijing 100083, China
d
CSIRO Energy, North Ryde, NSW 2113, Australia
e
Huaneng Clean Energy Research Institute, Beijing 102209, China
f
University of Newcastle, University Drive, Callaghan, NSW 2308, Australia
g
State Environmental Protection Key Laboratory of Integrated Surface Water-Groundwater Pollution Control, School of Environmental Science and Engineering, Southern
University of Science and Technology, Shenzhen 518055, China

ARTICLE INFO ABSTRACT

Keywords: CO2 mineralisation by industrial wastes is a promising option for mitigating carbon emissions safely and per-
CO2 sequestration manently with low material cost. But there is still absence of a detailed understanding on how fly ash properties
Mineral carbonation affect the carbonation reactions. To fill this knowledge gap, five coal combustion fly ashes, Beijing (BJ), Wuhai
Fly ash (WH), Hazelwood (HW), Yallourn (YA) and Loy Yang (LY) ashes, from China and Australia were selected for
Carbonation efficiency
carbonation studies. Experiments were performed in a batch reactor at 40 and 140 °C with 20 bar initial CO2
Mechanism
pressure, 200 g/L solid to liquid ratio, 450 rpm stirring rate to compare the carbonation performance of the five
fly ashes and the effect of fly ash properties on carbonation reactions. Then BJ, YA and HW ashes were then
selected for further study in a wide temperature range (40–220 °C) because of their higher CO2 sequestration
capacity than the other two ashes. Quantitative X-ray diffractometry (XRD) with Rietveld refinement was used to
characterize the crystalline and amorphous phases in fresh and carbonated fly ashes. Scanning electron mi-
croscopy (SEM) with energy dispersive spectrometry (EDS) were used to characterize morphological and ele-
mental properties of fresh and carbonated fly ash samples. Compared to LY and WH ashes, BJ, YA and HW ashes
displayed much higher CO2 sequestration capacity due to the higher fraction of reactive Ca/Mg-bearing crys-
talline phases, including lime (CaO) and portlandite (Ca(OH)2) in BJ ash, periclase (MgO) and srebrodolskite
(Ca2Fe2O5) in YA ash, and periclase and brucite (Mg(OH)2) in HW ash. Compared to YA and WH ashes, BJ ash
displayed faster kinetics of carbonation reactions because the reactant phases of BJ ash were mainly Ca-bearing
phases which have higher reactivity with CO2 than Mg-bearing phases. Also, particle size and morphological
analyses indicated that the reacted particles displayed a lower porosity and pore area than the fresh sample due
to the new precipitates not only depositing on the active surface, but also filling the pores of the fly ash particles,
which was responsible for the reduced kinetics with time.

1. Introduction in the short to medium term. CO2 sequestration by mineralisation can


permanently sequester CO2 emitted by fossil fuel combustion and is
Anthropogenic increases of carbon dioxide (CO2) concentration in considered a viable alternative to amine-based CO2 capture and geo-
the atmosphere have been widely accepted as the main cause of global logical storage. Accelerated natural rock weathering occurs when nat-
warming and climate change [1]. Fossil fuel combustion is the greatest ural alkaline minerals react with CO2 from the atmosphere in the pre-
contributor (~37%) of global CO2 emissions [2]. Given that natural sence of moisture/water to form stable carbonate minerals [3].
processes in the global carbon cycle cannot absorb all anthro- Unfortunately, the kinetics of naturally occurring CO2 mineralisation is
pogenically produced CO2 in the coming centuries [1], CO2 capture and very slow due to ambient conditions and low CO2 concentration
storage technologies are urgently required to mitigate global warming (0.03–0.06% by volume) in the atmosphere. Reaction kinetics can be


Corresponding author.
⁎⁎
Correspondence to: S. Zhao, Department of Environmental Sciences, Macquarie University, North Ryde, NSW 2109, Australia.
E-mail addresses: hai.yu@csiro.au (H. Yu), shuaifei.zhao@mq.edu.au (S. Zhao).

https://doi.org/10.1016/j.fuproc.2019.01.015
Received 16 November 2018; Received in revised form 22 January 2019; Accepted 31 January 2019
Available online 13 February 2019
0378-3820/ © 2019 Elsevier B.V. All rights reserved.
L. Ji et al. Fuel Processing Technology 188 (2019) 79–88

highly enhanced in the accelerated CO2 mineralisation technology, kinetics of CO2 mineralisation of fly ash within a wide temperature
where alkaline materials react with CO2 at elevated conditions and > range should also be investigated.
50% carbonation efficiency can be achieved in a few hours [4]. In this study, five fly ashes from China and Australia were selected for
Both alkaline natural minerals and industrial wastes can be used as carbonation. The experiments were first performed in a batch reactor at
feedstocks for CO2 mineralisation [5–8]. Natural minerals such as ser- 40 and 140 °C, 20 bar initial CO2 pressure, 200 g/L solid to liquid ratio,
pentine, olivine and wollastonite take advantages of the high calcium 450 rpm stirring rate to compare the carbonation characteristics of the
and magnesium contents and high worldwide abundance. Alkaline five fly ashes and investigate the effects of fly ash properties on carbo-
minerals near Earth's surface would be able to sequester all of the CO2 nation capacity. Then three ashes with better carbonation capacity were
produced from the recoverable fossil-fuel resources [9]. But the effec- selected for a further kinetics study in a wide range of temperature
tiveness of natural minerals-based CO2 mineralisation is limited by slow (40–220 °C). Quantitative X-ray diffraction (XRD) using Rietveld refine-
reaction rates between CO2 and Ca/Mg-silicates and huge energy pe- ment was used to characterize crystalline and amorphous phases in fresh
nalties of mineral pre-treatment, such as mechanical, chemical, or and carbonated fly ashes. Scanning electron microscopy (SEM) with
thermal pre-treatment. Recently, mineral carbonation using industrial energy dispersive spectrometry (EDS) was used to characterize mor-
by-products such as fly ash [10–15], cement kiln dust [16,17], steel- phological and elemental properties of fresh and carbonated fly ash
making slag [18], and mining wastes [19], have attracted attention due samples. The results can improve understanding of the governing pro-
to their low materials cost, high reactivity, no pre-treatment require- cesses and allow for better use of fly ash in CO2 sequestration.
ment, and availability near many CO2 emission sources.
Coal fly ash, a by-product of coal combustion power plants, has been 2. Materials and methods
considered a promising material for CO2 sequestration, because of its
high alkalinity and particularly large volume production worldwide 2.1. Materials
[20,21]. In 2010, the global generation of coal fly ash was approximately
780 Mt [22]. As a hazards waste, fly ash without proper disposal can Beijing (BJ) and Wuhai (WH) fly ashes used in this study were col-
pollute water and soil, disrupt ecological cycles and pose environmental lected from Huaneng Gaobeidian power plant in Beijing and Wuyun
hazards. Technologies have been developed to recycle this residue and power station in Wuhai, China, which were based on Chinese black coal.
use it as construction materials in cement, reclamation of low lying areas, Hazelwood (HW), Yallourn (YA) and Loy Yang (LY) fly ashes were col-
roads and embankments, mine filling and agricultural activities, but only lected from Hazelwood, Yallourn and Loy Yang thermal power plants in
53% of fly ash was utilised in the world in 2010 [23]. The annual gen- Australia respectively, which were based on Victorian lignite. Fly ash
eration of fly ash is increasing, and the unused fly ash is often stored in samples were dried for 24 h at 105 °C prior to testing. Elemental com-
waste piles or landfills. CO2 mineralisation by fly ash can not only reduce positions of fresh samples were determined using X-ray fluorescence
CO2 emissions, but also increase the stability of fly ash by reducing the spectrometry (XRF). Particle size distributions of the fresh samples were
alkaline contents through carbonation reactions, and thus reduce the measured by LS-POP (VI) laser particle analyzer. The summary of basic
potential hazard of fly ash disposal and expand further use. physicochemical properties of the fly ashes is shown in Table 1.
Previous studies have discussed the technical feasibility of direct
CO2 mineral carbonation with fly ash in both gas-solid and aqueous
2.2. Aqueous carbonation experiments in a batch reactor
routes [10,15]. Mineral carbonation processes conducted by the aqu-
eous route are more effective than those by dry gas-solid methods
Carbonation reactions were carried out in an autoclave Parr reactor
[24,25]. Extensive studies have investigated carbonation capacities and
made of Hastelloy alloy with an internal volume of 300 mL (Fig. 1). The
carbonation efficiencies of different fly ashes at various operation
internal volume of the vessel was calibrated prior to the experiments
conditions, such as temperature, CO2 pressure, solid to liquid ratio,
using N2 gas. For each carbonation experiment, 50 mL high-purity
stirring rate, and reaction time [12,20,24,25]. These studies indicated
water and 10 g fly ash were added to the autoclave. The slurry was
that carbonation capacity and efficiency of fly ash can be improved by
heated to a desired temperature (40, 60, 80, 100, 140, 180, and 220 °C)
optimising the operating parameters. However, there are limited stu-
using an oven specifically adapted to the reactor. Subsequently high
dies to understand how fly ash properties affect carbonation perfor-
purity CO2 gas was injected into the reactor from the gas cylinder and
mance. Carbonation performance can be significantly affected by fly
the initial pressure was 20 bar. Then mechanical stirring was turned on
ash chemical composition and mineralogy. Due to different raw coal
at 450 rpm and fly ash particles were immediately dispersed. This time
sources, coal combustion processes, fly ash collection methods, and
was recorded as the starting time of the carbonation reaction. Once the
implementation of environmental hazard mitigation techniques [26],
carbonation reaction started, the pressure drop inside the vessel as a
different fly ashes show different chemical properties and mineralogy.
function of time was monitored by a pressure sensor. The dissolution of
Even for fly ashes with similar elemental compositions, the distribution
CO2 in the solution during carbonation produced a global pressure drop
of elements in different minerals can be vary a lot. Calcium and mag-
nesium in different minerals displayed very different reactivities with
Table 1
CO2 in carbonation reactions [27,28]. Also, given that the reaction rate
Elemental composition of fly ash samples (presented as oxides).
of mineralisation is not fast enough for large-scale application under
current operating conditions, further studies of the reaction kinetics of Analyte Unit BJ YA LY WH HW
carbonation with fly ash are required to improve the understanding of
Chemical SiO2 % 42.8 8.9 42.8 40.9 5.8
reaction mechanisms thereby promoting carbonation reactions. The properties Al2O3 % 19.2 5.9 16.2 32.8 3.0
kinetics of CO2 mineralisation were also affected by operating para- Fe2O3 % 9.1 42.3 9.5 3.5 14.0
meters. Our previous study indicated that the stirring rate and the CaO % 16.4 9.4 3.6 13.4 32.4
MgO % 1.2 27.9 7.1 0.5 29.3
solid/liquid ratio had limited effects on the kinetics of mineral carbo-
TiO2 % 0.8 0.9 1.5 1.3 0.7
nation of the fly ash from Victorian lignite in a vessel reactor [29]. Na2O % 1.7 0.6 5.0 0.1 0.2
Initial high CO2 pressure enhanced the kinetics by increasing the CO2 K2 O % 1.5 0.2 0.3 0.6 0.2
concentration dissolved in water. However, the effect of temperature on P2O5 % 0.4 0.1 0.0 0.7 0.4
the kinetics of CO2 mineralisation of fly ash remained unclear. There- SO3 % 1.9 2.8 9.4 0.2 12.8
CO2 (titration) % 2.1 0.6 0.4 0.5 16.6
fore, a comprehensive understanding of the chemical and mineral
Theoretical carbonation capacity kg CO2/ 97.5 360 50.6 105 416
constituents of fly ashes is essential for sample selection and effective kg FA
enhancement of CO2 mineralisation. Effects of temperature on the

80
L. Ji et al. Fuel Processing Technology 188 (2019) 79–88

calculated as follows [21]:


Pcarbonation = Pglobal P blank solution (1)

The quantity of CO2 consumed by the carbonation process can be


calculated by:
Pcarbonation V 44
mCO2 (g CO2 /kg FA) =
ZRT mFA (2)

where V is the reactor volume occupied with gas, T is the temperature


of reaction, Z is the compressive factor of carbon dioxide collected from
property database of Aspen 8.0, which depends on the pressure and
temperature, and R is the gas constant (8.314 J/(mol K)) [16]. As the
pressure drop stopped and became stable, the vessel reactor was cooled
with cold water and depressurised. The suspension was immediately
filtered through a 0.2 μm membrane filter unit equipped with a vacuum
pump. The filter cake was dried overnight in an oven at 105 °C. The
fresh samples and some carbonated samples were selected to be ana-
lyzed by quantitative XRD and SEM-EDS to determine mineralogical
changes before and after carbonation reaction.
The theoretical carbonation capacity in fly ash can be calculated as
follows [30]:
44 44 44
Th_mCO2 = mCaO + mMgO × m SO3 mCO2,0
56 40 80 (3)

where Th_mCO2 (g-CO2/g-FA) is the theoretical carbonation capacity,


mCaO (g-CaO/g-FA), mSO3 (g-SO3/g-FA), and mMgO (g-MgO/g-FA) are
the weight fraction of CaO, SO3 and MgO in fresh samples respectively
and were calculated from XRF results. mCO2, 0 (g-CO2/g-FA) is the
Fig. 1. Schematic of the experimental system for mineral sequestration of CO2 weight fraction of CO2 in the fresh fly ash determined by acid titration.
by aqueous carbonation of fly ash in a stirred reactor. The carbonation efficiency can be calculated by Eq. (4).
mCO2
Table 2 [%] = × 100
Th_mCO2 (4)
Relative concentrations (wt%) of the phases identified in fresh fly ashes before
carbonation experiments.
Mineral Chemical formula BJ YA LY WH HW 2.3. Characterisation of fly ash samples using XRD and SEM-EDS

Quartz SiO2 3.3 3.3 38 2.6 1.3 The crystalline phases present in fresh and carbonated samples were
Microcline KAlSi3O8 – – 1.9 – –
Albite NaAlSi3O8 – 0.6 1.7 – –
determined by XRD analysis. Fly ash samples were ground in an agate
Diopside CaMgSi2O6 – 3.4 – – – mortar and pestle then packed into an aluminium holder. All samples
Augite (Ca,Na)(Mg,Fe,Al,Ti)(Si,Al)2O6 – – – – 2.2 were run on a Malvern Panalytical Empyrean X-ray diffractometer
Mullite Al6Si2O13 1.4 – – 35 – using CuKα radiation at 40 kV and 40 mA. Scans (2θ) were undertaken
Hematite Fe2O3 1.4 – 2.4 – –
from 2 to 90°, with a step interval of 0.02°. Phase identification was
Maghemite Fe2O3 – – 12 – –
Magnetite Fe3O4 1.8 – – 0.9 – performed using the Bruker Eva software package. Quantitative phase
Magnesioferrite MgFe2O4 – 44 – – – analysis was performed using Siroquant™ software following Rietveld
Spinel MgFe0.23Al1.77O4 – 3.5 9.5 – – refinement which generates a synthetic pattern which is then fitted to
Srebrodolskite Ca2Fe2O5 – 12 – – 3.0 the experimental data using a least squares minimisation procedure
Brownmillerite Ca2(Al,Fe)2O5 0.7 – – – –
Anhydrite CaSO4 2.1 2.8 – –
[31]. The advantage of this approach is that the complete diffraction
Gypsum CaSO4·2H2O – – – – 3.0 pattern of each mineral is used to derive the quantitative results rather
Bassanite CaSO4·0.5H2O – – 2.1 – 11 than relying upon several peaks for the determination. Although ap-
Portlandite Ca(OH)2 3.5 – – – – plication of Rietveld refinement usually requires that the phases be
Lime CaO 1.2 – – – –
crystalline to calculate the synthetic XRD pattern, Siroquant™ makes
Calcite CaCO3 – – – – 23
Brucite Mg(OH)2 – – – – 4.8 use of experimentally derived structural data (observed hkl files) for
Periclase MgO – 8.9 – – 3.2 amorphous or poorly crystalline phases and thus can be used to de-
Glauberite Na2Ca(SO4)2 – – 7.9 – – termine the amount of amorphous material present in the sample [32].
Thenardite Na2SO4 – – 3.0 – – Morphological investigations of fresh and carbonated fly ash sam-
Polyhalite K2Ca2Mg(SO4)4·2H2O – – 0.6 – –
Hydrotalcite Mg6Al2CO3(OH)16·4H2O – – – – 26
ples were performed with a SEM. For each test, fly ash samples were
Corundum Al2O3 – – – 3.7 – dispersed and mounted in epoxy resin. The epoxy disk was cured for
Amorphous 85 22 22 58 23 24 h at room temperature and standard pressure of N2. Once hardened,
the sample surface was ground flush using grinding paper. After
grinding, the sample was polished using micro Al2O3 powders. The
in the vessel reactor. To estimate the pressure-drop produced by car- polished sample was mounted with double-sided carbon tape on an
bonation, two complementary tests were conducted for each experi- aluminium stub, and then coated with a thin-layer of carbon. The SEM
ment. Firstly, the pressure drop by the dissolution of CO2 into pure was operated with 10 kV on the filament and a current density of
water was measured (Pblank solution). Secondly, the pressure reduction by 45 μA/cm2, with the electron beam directed at 90° to the specimen.
the dissolution of CO2 in the solution mixed with fly ash was measured Distributions of elements in the fresh and carbonated fly ash were de-
(Pglobal). The amount of CO2 consumed in the carbonation was termined using EDS in the SEM.

81
L. Ji et al. Fuel Processing Technology 188 (2019) 79–88

Fig. 2. X-ray diffractograms of fresh and corresponding carbonated ashes: (a) BJ ash (220 °C), (b) HW ash (220 °C) and (c) YA ash (180 °C).

82
L. Ji et al. Fuel Processing Technology 188 (2019) 79–88

Fig. 2. (continued)

Table 3 content of amorphous phases [12]. The initial CO2 contents of the
Particle size, surface area and porosity of fly ash samples. materials were 16.6, 2.1, 0.6, 0.5 and 0.4% for HW, BJ, YA, WH and LY
Items Unit BJ YA LY WH HW
ash respectively. Based on the chemical composition of these three fly
ashes, their theoretical capacities of CO2 sequestration were calculated
Volume mean diameter, D(4,3) μm 4.6 17.6 21.0 6.2 8.8 by Eq. (3), being 416, 360, 105, 97.5 and 50.6 g-CO2/kg-FA for HW,
Surface mean diameter, D(3,2) μm 3.0 15.6 16.8 4.5 6.0 YA, WH, BJ and LY ash, respectively.
D10 μm 1.7 10.8 10.5 2.7 3.5
D50 μm 4.4 17.7 20.7 5.9 8.4
D90 μm 7.9 24.1 31.9 9.9 14.6 3.1.2. Mineralogy
BET surface area m2/g 4.3 2.3 6.8 3.8 30.2 Table 2 shows the relative concentrations of minerals identified in
Total pore volume mm3/g 7.8 5.6 12.5 6.3 108 fresh ashes before carbonation experiments. BJ fresh ash was mainly
Average pore size nm 7.1 9.6 7.5 9.4 14.2
composed of amorphous phases (85%) while crystalline phases were
present in low concentrations (< 15% in total). Lime (CaO), portlandite
(Ca(OH)2), anhydrite (CaSO4) and brownmillerite (Ca2(Al,Fe)2O5) were
3. Results and discussion
the crystalline Ca-bearing phases, along with quartz (SiO2), mullite
(Al2Si6O13), hematite (Fe2O3) and magnetite (Fe3O4). The Ca fraction in
3.1. Physical and chemical properties of fly ash samples
crystalline phases was much smaller than the bulk fraction of Ca, in-
dicating that Ca was partially present in amorphous phases. There was
3.1.1. Elemental composition
no Mg-bearing mineral found in any crystalline phase, which indicated
Table 1 shows the elemental composition of the five fly ashes, which
all of magnesium was contained in amorphous phases. After carbona-
can help with preliminary sample selection and the phase peak iden-
tion reactions, two newly formed carbonates were identified in the
tification in XRD analysis. Since CaO and MgO are the ideal feedstocks
carbonated sample: calcite (CaCO3) and aragonite (CaCO3) (Fig. 2(a)).
for CO2 mineral carbonation, the fly ashes obtained for this study
Lime and portlandite were not detected in the carbonated sample, in-
generally exhibited a high fraction of CaO and MgO indicating their
dicating that they reacted with CO2 and produced calcite or aragonite
potential for CO2 sequestration. HW ash had the highest CaO fraction
during carbonation reactions. The peak intensity of anhydrite, and
(32.4%), followed by BJ ash (16.4%), WH (13.4%), YA ash (9.4%), and
brownmillerite reduced after carbonation, indicating that they also
LY ash (3.6%). HW ash exhibited the highest MgO content (29.3%),
reacted with CO2. Thus, the involved phases were Ca-bearing phases:
followed by YA ash (27.9%), LY ash (7.1%), BJ ash (1.2%), and WH ash
lime, portlandite, anhydrite and brownmillerite.
(0.5%). The high SO3 content of HW ash (12.8%) and LY ash (9.4%)
The fresh HW ash exhibited a much lower fraction of amorphous
suggested their calcium might be present as calcium sulfate which
phase (23%) compared to BJ fly ash (85%), and a more complex
would result in low calcium amounts available for carbonation reac-
composition of crystalline phases (Table 2). Calcite, basanite (Ca-
tions [12]. The high SiO2 and Al2O3 fractions of BJ ash (42.8% and
SO4·0.5H2O), augite ((Ca,Na)(Mg,Fe,Al,Ti)(Si,Al)2O6), gypsum (Ca-
19.2% respectively), LY ash (42.8% and 16.2% respectively) and WH
SO4·2H2O) and srebrodolskite (Ca2Fe2O5), with weight fractions of
ash (40.9% and 32.8% respectively) indicated their probable high
23%, 11%, 3.0%, 3.0%, and 2.2%, respectively, were the crystalline Ca-

83
L. Ji et al. Fuel Processing Technology 188 (2019) 79–88

peak intensity of several Mg-bearing phases reduced, including diop-


side, spinel, srebrodolskite and periclase (Fig. 2(c)), indicating that they
were consumed in carbonation reactions.
The fresh LY ash was found to be mainly composed of amorphous
phases (22%) and quartz (SiO2, 38%), in addition to several mixed
oxide phases: hematite, spinel, bassanite, glauberite (Na2Ca(SO4)2) and
thenardite (Na2SO4). The fresh WH ash was also dominated by amor-
phous phases (58%) and mullite (35%), in addition to several other
crystalline phases with low weight fractions: corundum (Al2O3, 3.7%),
quartz (2.6%), and magnetite (0.9%). LY and WH ashes might have low
carbonation capacities as there was no appreciable faction of CaO/MgO
phases identified in these two ashes.

3.1.3. Particle size, surface area and porosity


Table 3 shows the particle size, surface area and porosity of the five
ashes. Compared to the other four ashes, BJ ash displayed the largest
proportion of fine particles, with a D50 of 4.4 μm and a D90 of 7.9 μm.
The small particle size probably provided better mass transfer for Ca/
Mg leaching in carbonation reactions [33]. LY ash contained the largest
proportions of coarse particles with a D90 of 31.9 μm. As the particles of
these five ashes were very fine, the pretreatment of grinding was not
needed prior to mineral carbonation. A previous study indicated that
the grinding can be avoided when the particle size of the feedstock
was < 100 μm [34]. HW ash exhibited much larger BET surface area
than other four ashes, and thus it can provide a larger interaction
surface between dissolved CO2 and the ash particles which benefits
reaction kinetics.
Fig. 3 shows the particle size distribution of three selected ashes
before and after carbonation. The fresh BJ, HW and YA ash displayed a
unimodal trend and consisted of one major particle population with
grain sizes in the range of 3–17, 5–20 and 13–30 μm, respectively. Size
distribution curves of the carbonated samples (B3, H4 and Y3) in-
dicated the increased fraction of large particles after carbonation re-
actions. Similar observations have been reported [35]. Compared with
B3, H4 and Y3, a further increase of particle size was observed in B1, H1
and Y1. As the carbonation efficiency of samples B1, H1 and Y1 (Fig. 7)
was larger than that of samples B3, H4 and Y3, the particle size of
carbonated samples was strongly affected by carbonation reactions and
was related to carbonation efficiency. The increased particle size might
be attributed to the combined effects of the dissolution of fine particles
and the newly formed precipitates that covered the un-reacted particles
[33–35]. The surface area and total pore volume of the carbonated
samples decreased significantly after carbonation, indicating that the
newly formed precipitates also filled some pores [36–38]. The newly
formed precipitate on the active surface and infilled pores hindered the
mass transfer of carbonation reactions, which was partly responsible for
Fig. 3. Particle size distribution of fresh and selected carbonated fly ashes: (a)
the reduced kinetics with time [38].
BJ ash, (b) HW ash and (c) YA ash.

3.1.4. Morphology
bearing phases identified, while the Mg-bearing phases included brucite Fig. 4(a–f) are the SEM images of the fresh and carbonated ash
(Mg(OH)2), periclase and pyroaurite (Mg6 Fe2CO3(OH)16·4H2O), with samples. Compared to fresh samples (Fig. 4(a), (c) and (e)), the SEM
weight fractions of 4.8%, 3.2% and 26%, respectively. After carbona- images of the carbonated samples (Fig. 4(b), (d) and (f)) show that the
tion, the significantly increased peak intensity of calcite suggests the newly formed Ca-carbonate crystals covered the unreacted particles
formation of calcite in the carbonation reaction (Fig. 2(b)). Brucite and resulting in “grouped” particles, and that some small particles dis-
periclase participated in carbonation reactions as their peak dis- appeared after carbonation reactions. This observation confirmed the
appeared after carbonation. Srebrodolskite was another reactant of increased particle sizes of carbonated samples (Fig. 3). With a semi-
carbonation reaction as its peak intensity reduced after carbonation. quantitative chemical analysis by EDS, it is possible to identify the
The peak intensity of gypsum and basanite also reduced after the re- elemental composition of the precipitates and the unreacted ash par-
actions. Thus, the reacted phases of HW ash include both Ca-bearing ticles, and to roughly assess which phases were present in the fly ash
phases (srebrodolskite, gypsum and basanite) and Mg-bearing phases samples [39], and thus provide confirmation for the XRD results. The
(brucite and periclase). EDS analysis (Supporting Information Fig. S1) shows that the pre-
The crystalline component of YA fresh ash was dominated by Ca- cipitates surrounding the unreacted BJ ash particles (Fig. 4b) were
and Mg-bearing phases (Table 2), including magnesioferrite (MgFe2O4), composed of a Ca−C−O phase, and the unreacted particles were
srebrodolskite (Ca2Fe2O5), periclase (MgO), spinel (MgFe0.23Al1.77O4), mainly composed of a Ca−Al−Si−O phase. The elemental composi-
diopside (CaMgSi2O6), and anhydrite and with weight fractions of 44%, tion of the newly formed precipitate was consistent with the Ca-calcite
12%, 8.9%, 3.5%, 3.4% and 2.8%, respectively. After carbonation, the phase detected using XRD (Fig. 2a). For HW ash, the newly formed

84
L. Ji et al. Fuel Processing Technology 188 (2019) 79–88

Fig. 4. Cross-sections of the fly ash samples using SEM-EDS: (a) fresh BJ ash, (b) carbonated BJ ash, (c) fresh HW ash, (d) carbonated HW ash, (e) fresh YA ash, (f)
carbonated YA ash.

Fig. 5. Performance comparison of the five fly ashes in carbonation reaction at


temperature of 40 and 140 °C, initial CO2 pressure of 20 bar, solid to liquid ratio
of 200 g/L, stirring rate of 450 rpm. Fig. 6. Thermodynamic equilibrium calculation for reactions of different Ca-/
Mg-bearing minerals with CO2.

85
L. Ji et al. Fuel Processing Technology 188 (2019) 79–88

identified as anhydrite and diopside, respectively. The presence of a


Mg−C−O phase indicated formation of Mg−carbonates after carbo-
nation reactions.

3.2. Carbonation performance comparison of the selected fly ashes

3.2.1. Capacity comparison of the five selected fly ash samples


The CO2 sequestration capacity and carbonation efficiency of the
five fly ashes were measured at 40 and 140 °C, initial CO2 pressure of
20 bar, solid to liquid ratio of 200 g/L, stirring rate of 450 rpm.
Compared to other ashes, HW ash displayed the highest carbonation
capacity at 40 and 140 °C, being 103.0 and 102.0 g-CO2/kg-FA re-
spectively (Fig. 5). The highest carbonation capacity of HW ash was due
to its largest amount of Ca/Mg-bearing phases available for carbona-
tion. But the carbonation efficiency was low (about 24%) at 40 °C,
which indicated that most of the Ca-/Mg-bearing phases cannot react
with CO2. Increasing the reaction temperature to 140 °C did not pro-
mote carbonation efficiency. Similar observations have been reported
previously [39], and the reason might be that un-reacted phases were
stable in this temperature range (40–140 °C). Although the carbonation
capacity of BJ ash was much smaller than HW ash, being 35.9 and
43.2 g-CO2/kg-FA at 40 and 140 °C respectively, BJ ash displayed the
highest carbonation efficiency, of 36.8% and 44.4% at 40 and 140 °C
respectively (Fig. 5). Given that the reactants of BJ ash were mainly
lime and portlandite (Fig. 2a) while the reacted phases of HW ash in-
cluded both Ca-bearing phases (srebrodolskite, gypsum and basanite)
and Mg-bearing phases (brucite and periclase), the Ca-bearing phases
might be more reactive with CO2. Lime and portlandite phases of BJ ash
were almost completely converted to calcite after carbonations
(Fig. 2a). Different from BJ and WH ash, YA ash exhibited a significant
increase of carbonation capacity from 33.3 to 93.1 g-CO2/kg-FA when
the temperature was elevated from 40 to 140 °C, probably because the
elevated temperature made more phases reactive with CO2. Fig. 6
shows the Gibbs free energy change (ΔG) of various Ca/Mg-oxide/hy-
droxides reacting with CO2. Mg oxide/hydroxide displayed higher ΔG
than Ca hydroxide. YA presented low weight fraction of Ca-bearing
phases but high weight fraction of Mg-bearing phases (Table 2). This
explains the obvious improved carbonation capacity of YA ash at
140 °C. Compared to the above three ashes, LY and WH ash displayed
much lower carbonation capacity at the same conditions, which was
attributed to the small fraction of CaO/MgO phases identified in their
fresh samples.

3.2.2. Performance comparison of BJ, HW and YA ash within a wide range


temperature
Fig. 7a–c present the carbonation capacity and efficiency of BJ, HW
and YA ash in carbonation reactions as a function of time at different
temperatures (40, 60, 80, 100, 140, 180, and 220 °C, respectively). The
three fly ashes exhibited the same trend of the carbonation rate de-
creasing as the reaction time elapsed. The carbonation capacity/effi-
ciency increased rapidly in the first 20 min and reached a maximum
value after 120 min, consistent with the literature [40,41]. More than
80% of the maximum carbonation efficiency was achieved in 20 min.
Fig. 7. Model fitting results of carbonation experiments of (a) BJ ash, (b) HW This might be attributed to the exhaustion of the reactant on the surface
ash and (c) YA ash at temperatures of 40, 60, 80, 100, 140, 180 and 220 °C, of the ash particles, and the formation of a precipitate layer at the early
initial CO2 pressure of 20 bar, solid to liquid ratio of 200 g/L, stirring rate of stage of carbonation, which hindered diffusion of the reactant inside the
450 rpm.
particles [36,37]. Compared to YA and WH ashes, BJ ash displayed
faster carbonation kinetics because the reactant phases of BJ ash were
Ca−Mg−C−O phases (Fig. 4d) likely present as Ca/Mg-carbonates. mainly Ca-bearing phases which have higher reactivity (Fig. 6) with
The Ca−C−O and Mg−C−O phases (Fig. 4d) were consistent with Ca- CO2 than Mg-bearing phases [34,35].
calcite and Mg-calcite (Fig. 2b). The precipitates contained not only The three fly ash displayed different trend of carbonation efficiency
calcium carbonates, but also magnesium carbonates. In the fresh YA ash within the temperature range used in this study. Specifically, the car-
(Fig. 4e), the bright Mg−Fe−O and Fe−O rich phases were common. bonation efficiency of BJ ash within 120 min at different temperature
These phases exclusively featured angular boundaries, and thus were decreased in the order: 220 °C > 180 °C > 140 °C > 40 °C >
identified as crystalline magnesioferrite [39]. In the carbonated YA 60 °C > 100 °C > 80 °C. When the temperature increased from 40 to
sample (Fig. 4f), Ca−S−O and Ca−Mg−Si−O rich phases were 220 °C, carbonation efficiency decreased first and then increased, due to

86
L. Ji et al. Fuel Processing Technology 188 (2019) 79–88

Fig. 8. Schematic of major material flows for mineral carbonation of coal-fired power generation based on typical Australian conditions.
(Modified after Geoffrey et al. [43]).

the complex effects of temperature on carbonation. Elevated tempera- ash or 19 YA ash. Considering that < 0.1 t fly ash is normally produced
tures can increase reaction rates by improving mass transfer rates, by the combustion of 1 t coal, coal fly ash alone cannot be used to
promoting thermal motion of molecules and increasing their average significantly reduce CO2 emissions from coal fired power plants. De-
kinetic energy which helped speed up carbonation [35]. Raising reac- spite the low capacity, CO2 mineralisation by coal fly ash is still at-
tion temperature also reduced the solubility of carbon dioxide in the tracting wide attention because of the huge volumes of annual fly ash
solution [17] and the precipitation of the carbonate product [29,42]. production (approximately 780 Mt world-wide, and 11 Mt in Australia)
Thus, the kinetic results of BJ ash indicated that the reduced CO2 so- and can be a viable alternative to amine-based CO2 capture and geo-
lubility had a larger effect on carbonation resulting in a low carbona- logical storage. Also, the carbonated ash has great potential to be used
tion efficiency in the temperature range 40 to 80 °C, while the enhanced as construction materials, such as cement, concrete, aggregate etc.
mass transfer rate at elevated temperatures might have a larger effect in (Fig. 8). The changes in physico-chemical properties of carbonated fly
the temperature range 80 to 220 °C. Similar to BJ ash, the carbonation ash were found to be beneficial to sequential uses in cement and con-
efficiency of HW ash decreased first from 40 to 80 °C and then obviously crete [11]. After the carbonation reaction, the potential for heavy metal
increased when the temperature increased from 80 to 220 °C. However, leaching and uncontrollable expansion can be eliminated [44,45]. After
the carbonation efficiency increase of HW ash is much larger than BJ carbonation, both the CaO, Ca(OH)2 etc. contents significantly de-
ash when temperature was elevated from 80 to 220 °C. In addition, the creased and were converted into carbonates [12]. The surface of the
carbonation efficiency of YA ash decreased from 40 to 60 °C and then fresh fly ash was normally smooth, while the surface of carbonated fly
increased thereafter which is different from BJ and HW ashes. Con- ash particles was covered by rhombohedral (i.e., calcite) and/or needle
sidering the same operating parameters used in the carbonation ex- like (i.e., aragonite) crystals resulting in arcuately shaped and a larger
periments, the different carbonation efficiency trends of the three ashes surface area. The larger surface area means greater mechanical strength
within the used temperature range results from their different miner- of cement. In addition, the formation of the fine CaCO3 precipitates can
alogical properties. The detailed investigation of the mineral changes provide a favourable surface for nucleation and growth of hydration
before and after carbonation reactions in Section 3.1.2 could thus help products in cement/concrete, which is beneficial to the strength de-
explain this phenomenon, as well as the ΔG of the reacted minerals. The velopment of blended cement. Pan et al. [11] evaluated the workability,
main reacted minerals of BJ ash, Ca-bearing minerals, owns very low strength development, and durability of the blended cement with dif-
ΔG which means high reactivity with CO2, and thus leads to its fast ferent substitution ratios (i.e., 10%, 15%, and 20%) of carbonated fly
kinetics and large carbonation efficiency in low temperature. For YA ash. The results indicated that cement with carbonated fly ash exhibited
ash, it was difficult for srebrodolskite, spinel and diopside phases to superior initial compressive strength and durability compared to ce-
react with CO2 at low temperature due to their low reactivity. ment with fresh fly ash.

3.3. Assessment of the CO2 sequestration capacity by coal fly ash 4. Conclusions

The results of this study reveal that the maximum CO2 sequestration In the present study, the carbonation capacity and carbonation ef-
capacity of HW, BJ and YA ashes are 132, 58 and 125 g-CO2/kg fly ash, ficiency of BJ, WH, HW, YA and LY ashes were systematically in-
respectively. This capacity seems smaller compared to that of some vestigated. Compared to LY and WH ashes, BJ, YA and HW ashes dis-
other industrial wastes such as steel slag. Previous study [43] indicated played much higher carbonation capacity due to the higher faction of
that 2.4 t CO2 was normally regenerated by the combustion of 1 t coal reactive Ca/Mg-bearing crystalline phases, including lime and por-
normally. Sequestrating these CO2 would require 18 t HW ash, 41 t BJ tlandite in BJ ash, periclase and srebrodolskite in YA ash, and periclase

87
L. Ji et al. Fuel Processing Technology 188 (2019) 79–88

and brucite in HW ash. Compared to YA and WH ash, BJ ash displayed 692–705.


faster kinetics of carbonation because the reactant phases of BJ ash [18] R. Baciocchi, et al., Accelerated Carbonation of Steel Slags Using CO2 Diluted
Sources: CO2 Uptakes and Energy Requirements, Front. Energy Res. 3 (2015) 1–10.
were mainly Ca-bearing phases which were more reactive with CO2 [19] A.L. Harrison, et al., Accelerated carbonation of brucite in mine tailings for carbon
than the Mg-bearing phases in YA and WH ashes. Fast kinetics can be sequestration, Environ. Sci. Technol. 47 (1) (2013) 126–134.
achieved by selecting suitable fly ash samples as the feedstock. The [20] W. Liu, et al., CO2 sequestration by direct gas–solid carbonation of fly ash with
steam addition, J. Clean. Prod. 178 (2018) 98–107.
kinetics of mineral carbonation can be effectively enhanced by ele- [21] G. Montes-Hernandez, et al., Mineral sequestration of CO2 by aqueous carbonation
vating temperature, especially for fly ash samples containing Mg- of coal combustion fly-ash, J. Hazard. Mater. 161 (2–3) (2009) 1347–1354.
bearing crystalline phases. Particle size analyses and morphological [22] C. Heidrich, et al., Coal combustion products: a global perspective, World of Coal
Ash (WOCA) Conference, April 22–25, Lexington, KY, America, 2013.
investigations indicated that the reacted particles displayed a lower [23] Z.T. Yao, et al., A comprehensive review on the applications of coal fly ash, Earth
porosity and pore area than fresh samples due to the newly formed Sci. Rev. 141 (2015) 105–121.
precipitate not only having deposited on the active surface, but also [24] C.W. Noack, et al., Comparison of alkaline industrial wastes for aqueous mineral
carbon sequestration through a parallel reactivity study, Waste Manag. 34 (10)
filling the pores of the fly ash particles, which was responsible for the
(2014) 1815–1822.
reduced kinetics with time. [25] R.R. Tamilselvi Dananjayan, et al., Direct mineral carbonation of coal fly ash for
Supplementary data to this article can be found online at https:// CO2 sequestration, J. Clean. Prod. 112 (2016) 4173–4182.
doi.org/10.1016/j.fuproc.2019.01.015. [26] Y. Li, X. Gao, H. Wu, Further Investigation into the formation mechanism of ash
cenospheres from an Australian coal-fired power station, Energy Fuel 27 (2013)
811–815.
Acknowledgment [27] C. Ward, D. French, Determination of glass content and estimation of glass com-
position in fly ash using quantitative X-ray diffractometry, Fuel 85 (2006)
2268–2277.
Long Ji thanks Macquarie University for the Cotutelle-iMQRES [28] S. Vassilev, Phase-mineral and chemical composition of coal fly ashes as a basis for
scholarship, and to CSIRO Energy and Huaneng Clean Energy Research their multicomponent utilization. 1. Characterization of feed coals and fly ashes,
Institute for the opportunity to work in their laboratories and use their Fuel 82 (14) (2003) 1793–1811.
[29] L. Ji, et al., Insights into carbonation kinetics of fly ash from Victorian lignite for
resources. CO2 sequestration, Energy Fuel 32 (2018) 4569–4578.
[30] E.E. Chang, et al., Accelerated carbonation using municipal solid waste incinerator
References bottom ash and cold-rolling wastewater: performance evaluation and reaction ki-
netics, Waste Manag. 43 (2015) 283–292.
[31] H.B. Vuthaluru, D.H. French, Mineralogical investigations into clinker formation
[1] M. Bui, et al., Carbon capture and storage (CCS): the way forward, Energy Environ. and variations in deposit characteristics with time in a large-scale pc-fired boiler,
Sci. 11 (5) (2018) 1062–1176. Fuel 150 (2015) 184–190.
[2] Y. Tan, et al., Property impacts on Carbon Capture and Storage (CCS) processes: a [32] C.R. Ward, et al., Quantitative mineralogy of sandstones by X-ray diffractometry
review, Energy Convers. Manag. 118 (2016) 204–222. and normative analysis, J. Sediment. Res. 69 (5) (1999) 1050–1062.
[3] A. Sanna, et al., A review of mineral carbonation technologies to sequester CO2, [33] A. Polettini, et al., Carbon sequestration through accelerated carbonation of BOF
Chem. Soc. Rev. 43 (23) (2014) 8049–8080. slag: influence of particle size characteristics, Chem. Eng. J. 298 (2016) 26–35.
[4] E.R. Bobicki, et al., Carbon capture and storage using alkaline industrial wastes, [34] R. Baciocchi, et al., Thin-film versus slurry-phase carbonation of steel slag: CO2
Prog. Energy Combust. Sci. 38 (2) (2012) 302–320. uptake and effects on mineralogy, J. Hazard. Mater. 283 (0) (2015) 302–313.
[5] A. Sanna, et al., Waste materials for carbon capture and storage by mineralisation [35] A. Polettini, et al., CO2 sequestration through aqueous accelerated carbonation of
(CCSM) – a UK perspective, Appl. Energy 99 (2012) 545–554. BOF slag: a factorial study of parameters effects, J. Environ. Manag. 167 (2016)
[6] X. Wang, M.M. Maroto-Valer, Optimization of carbon dioxide capture and storage 185–195.
with mineralisation using recyclable ammonium salts, Energy 51 (2013) 431–438. [36] S.-Y. Pan, et al., Kinetics of carbonation reaction of basic oxygen furnace slags in a
[7] X. Wang, M.M. Maroto-Valer, Dissolution of serpentine using recyclable ammonium rotating packed bed using the surface coverage model: maximization of carbonation
salts for CO2 mineral carbonation, Fuel 90 (3) (2011) 1229–1237. conversion, Appl. Energy 113 (0) (2014) 267–276.
[8] X. Wang, M.M. Maroto-Valer, Integration of CO2 capture and mineral carbonation [37] S.Y. Pan, et al., Systematic approach to determination of maximum achievable
by using recyclable ammonium salts, ChemSusChem 4 (9) (2011) 1291–1300. capture capacity via leaching and carbonation processes for alkaline steelmaking
[9] R. Zevenhoven, et al., CO2 mineral sequestration: developments toward large-scale wastes in a rotating packed bed, Environ. Sci. Technol. 47 (23) (2013)
application, Greenhouse Gases Sci. Technol. 1 (1) (2011) 48–57. 13677–13685.
[10] S.L. Pei, et al., Environmental benefit assessment for the carbonation process of [38] D. Medas, et al., Accelerated carbonation by cement kiln dust in aqueous slurries:
petroleum coke fly ash in a rotating packed bed, Environ. Sci. Technol. 51 (18) chemical and mineralogical investigation, Greenhouse Gases Sci. Technol. 0 (2017)
(2017) 10674–10681. 0–15.
[11] S.-Y. Pan, et al., Integrated CO2 fixation, waste stabilization, and product utilization [39] R.T. Chancey, et al., Comprehensive phase characterization of crystalline and
via high-gravity carbonation process exemplified by circular fluidized bed fly ash, amorphous phases of a Class F fly ash, Cem. Concr. Res. 40 (1) (2010) 146–156.
ACS Sustain. Chem. Eng. 4 (6) (2016) 3045–3052. [40] Y. Sun, et al., Sequestration of carbon dioxide by indirect mineralization using
[12] L. Ji, et al., CO2 sequestration by direct mineralisation using fly ash from Chinese Victorian brown coal fly ash, J. Hazard. Mater. 209–210 (2012) 458–466.
Shenfu coal, Fuel Process. Technol. 156 (2017) 429–437. [41] N.L. Ukwattage, et al., A laboratory-scale study of the aqueous mineral carbonation
[13] L. Ji, et al., Integrated absorption–mineralisation for energy-efficient CO2 seques- of coal fly ash for CO2 sequestration, J. Clean. Prod. 103 (2014) 665–674.
tration: Reaction mechanism and feasibility of using fly ash as a feedstock, Chem. [42] W. Wang, et al., CO2 fixation in Ca2+/Mg2+ rich aqueous solutions through en-
Eng. J. 352 (2018) 151–162. hanced carbonate precipitation, Ind. Eng. Chem. Res. 50 (13) (2011) 8333–8339.
[14] L. Ji, et al., Integrated absorption-mineralisation for low-energy CO2 capture and [43] G.F. Brent, et al., Mineral carbonation as the core of an industrial symbiosis for
sequestration, Appl. Energy 225 (2018) 356–366. energy-intensive minerals conversion, 16 (1) (2012) 94–104.
[15] L. He, et al., A novel method for CO2 sequestration via indirect carbonation of coal [44] M.N. Grace, et al., Geochemical modeling of brine remediation using accelerated
fly ash, Ind. Eng. Chem. Res. 52 (43) (2013) 15138–15145. carbonation of fly ash, Desalin. Water Treat. 57 (11) (2015) 4853–4863.
[16] T. Wang, et al., Accelerated mineral carbonation curing of cement paste for CO2 [45] M.N. Grace, Statistical testing of input factors in the carbonation of brine impacted
sequestration and enhanced properties of blended calcium silicate, Chem. Eng. J. fly ash, J. Environ. Sci. Health A Tox. Hazard. Subst. Environ. Eng. 47 (2) (2012)
323 (2017) 320–329. 245–259.
[17] D. Medas, et al., Accelerated carbonation by cement kiln dust in aqueous slurries:
chemical and mineralogical investigation, Greenhouse Gases Sci. Technol. 7 (2017)

88

You might also like