You are on page 1of 10

The Journal of Supercritical Fluids 171 (2021) 105191

Contents lists available at ScienceDirect

The Journal of Supercritical Fluids


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

Pressurized in situ X-ray diffraction insights into super/subcritical


carbonation reaction pathways of steelmaking slags and constituent silicate ]]
]]]]]]
]]

minerals
Ye Eun Chaia,1, Quin R.S. Millerb,2, H. Todd Schaefb,3, Dushyant Barpagac,4,
Reza Bakhshoodehd,5, Marius Bodore,6, Tom Van Gervenf,7, Rafael M. Santosa, ,8

a
School of Engineering, University of Guelph, Guelph, Ontario N1G 2W1, Canada
b
Physical and Computational Sciences Directorate, Pacific Northwest National Laboratory, Richland, WA 99352, USA
c
Energy and Environment Directorate, Pacific Northwest National Laboratory, Richland, WA 99352, USA
d
Department of Civil, Environmental and Mining Engineering, The University of Western Australia, Perth, Western Australia 6009, Australia
e
Department of Materials and Environmental Engineering, "Dunarea de Jos" University of Galati, Galaţi 800201, Romania
f
Department of Chemical Engineering, KU Leuven, Leuven 3001, Belgium

H IGHL IGH TS GRA PHICA L A B ST RAC T

• Slags (AOD, CC) and six synthetic si-


licates are carbonated in wet CO2(g)
and scCO2.
• In situ high-pressure XRD analysis al-
lows tracking the carbonation path
and rate.
• Ca- and Mg-carbonates that form are
time- and silicate-dependent.
• Real-time monitoring of mineralogy
enables detecting transient and un-
stable phases.
• Varying carbonation behavior of sili-
cates motivates better slag chemistry
control.

A R T I C L E I N F O A B S T R A C T

Keywords: This study explores mineral carbonation of industrial stainless steelmaking slags and relevant synthetic con-
Accelerated mineral carbonation stituent minerals via in situ pressurized X-ray diffraction, to clarify carbonation reaction pathways and efficiency
Time-resolved XRD for carbon storage and waste valorization. The primary mineral phases of Argon Oxygen Decarburization (AOD)
Silicate and carbonate mineralogy and Continuous Casting (CC) slags, namely, åkermanite, bredigite, cuspidine, merwinite, and β- and γ-C2S, were
Carbon sequestration
reacted in a custom-built beryllium-capped XRD reactor filled with either water-saturated (wet) subcritical


Corresponding author.
E-mail address: santosr@uoguelph.ca (RM. Santos).
1
0000-0001-6923-7979
2
0000-0003-3009-9702
3
0000-0002-4546-3979
4
0000-0003-2271-6213
5
0000-0001-5921-3957
6
0000-0002-8056-4556
7
0000-0003-2051-5696
8
0000-0002-8368-8618

https://doi.org/10.1016/j.supflu.2021.105191
Received 22 September 2020; Received in revised form 10 January 2021; Accepted 4 February 2021
Available online 10 February 2021
0896-8446/© 2021 Elsevier B.V. All rights reserved.
Y.E. Chai, Q.R.S. Miller, H.T. Schaef et al. The Journal of Supercritical Fluids 171 (2021) 105191

Supercritical CO2 CO2(g), or wet supercritical CO2(SC), in a series of carbonation experiments. Formation of calcite and aragonite
Slag valorization was observed for most Ca-bearing minerals, transient precipitation of metastable vaterite was observed, while
carbonation of AOD and CC slags in CO2(SC) resulted in hydrated crystalline calcium carbonates and indirect
evidence of amorphous carbonate. Results obtained suggest that low-pressure subcritical carbonation routes are
attractive for standalone slag carbonation processes, while high-pressure supercritical carbonation routes are
amiable to symbiotic integration with CO2(SC)-generating green technologies.

1. Introduction mineral carbonation [15]. While low compared to global CO2 emis-
sions, such reduction potential of steelmaking slag will be significant
Global warming has stimulated renewed interest in issues relating to for an individual steel mill.
carbon capture and storage to reduce CO2 emissions from fossil-fuel Due to its suitable chemistry and mineralogy, BF slag, once ex-
power stations and other industrial facilities. Since carbon dioxide tensively stockpiled, has found utilization in the construction domain,
(CO2) is the most significant anthropogenic greenhouse gas [1], its re- as an additive in lightweight mortars, as a component of concrete
duction is crucial for climate change mitigation. Mineral carbonation is mixtures, as aggregates in road building, or as a raw material for the
a CO2 sequestration method that can safely store CO2 as a solid com- preparation of ceramic glass, silica gel, ceramic tiles and bricks
pound with very low risk for leakage. In mineral carbonation, CO2 is [16–18]. Similar aims are intended for steelmaking slags, including BOF
fixed by promoting calcium or magnesium oxides in silicate-based (also known as converter slag), EAF, Argon Oxygen Decarburization
materials to react with CO2 and form stable solid carbonates [2]. This (AOD) (also known as Linz-Donawitz slag), and Continuous Casting
approach has a larger capacity for CO2 storage than any of the other (CC) slags. Yet, various reasons, such as high basicity and the pro-
CO2 storage method because suitable silicate mineral resources, con- pensity for heavy metal leaching and pulverization [17,19–23], have
taining calcium or magnesium oxides, are abundant [3]. In addition, made the reutilization of these materials challenging. In order to im-
mineral carbonation allows carbon utilization, in addition to capture prove the negative properties of steelmaking slags and enable greater
and sequestration, when the carbonated products are used in large-scale opportunities for recycling, steelmaking slags have been the subject of
applications, such as in construction materials [4–6]. mineral carbonation research in recent years, with the purpose of CO2
While there have been commercial implementations of mineral sequestration, due to the geochemical stability of alkaline earth metal
carbonation in recent years, it remains challenging to implement mi- carbonates, and for waste valorization [17,24–27].
neral carbonation at a large scale. Bourgeois et al. [7] argue that suc- Process limitations, including high energy intensity, low reaction
cessful mineral carbonation processes must become part of a viable conversion, and slow reaction kinetics, have prevented mineral carbo-
economic framework that uses resources available in a given region to nation of waste materials from being widely applied. These barriers are
produce commercializable materials that meet identified market needs. caused by inefficient or nonoptimal processing and reactor technolo-
This also entails consideration of economic, societal, and environmental gies, morphological limitations (e.g. shrinking core model), and mi-
factors, beyond the technical feasibility of mineral carbonation pro- neralogical aspects inherent to the materials. While the first two mo-
cesses. Notably, mineral carbonation applied to alkaline waste mate- tives have been well reported in literature [17], limited attention has
rials has the potential to address these additional factors. Waste va- been paid to mineralogical susceptibility of individual minerals
lorization aims to generate added value, by commercialization of [28–30]. Alkaline waste materials are typically composed of several
upcycled products or avoidance of waste disposal, storage or treatment mineral phases that may or may not be susceptible to mineral carbo-
costs, while also reducing the environmental impact or risk of industrial nation, and which, if reactive to CO2, may exhibit varying degrees of
operations, and addressing societal concerns on the sustainability of the carbonation kinetics and influence the material’s basicity. The forma-
energy- and resource-intensive industries that generate these wastes. tion and character of passivating layers (e.g., thickness, porosity) can
Renforth [8] has recently estimated that 7 billion tonnes of alkaline also be affected by the relative solubility of the minerals [31,32].
waste materials are annually produced globally, possessing a CO2 sto- We have previously studied the carbonation of AOD and CC slags,
rage potential of 2.9–8.5 billion tonnes per year by 2100. Renforth [8] which have powdery morphology that provides sufficient surface area
identified the top five industrial waste materials in terms of carbonation for the active reaction during direct aqueous carbonation [33]. By
potential, from highest to lowest, as follows (in kg CO2 per tonne units): utilizing two carbonation methodologies, unpressurized thin-film car-
blast furnace (BF) slag (413 ± 13); basic oxygen furnace (BOF) slag bonation and pressurized slurry carbonation, carbonation of CC slag
(402 ± 17); electric arc furnace (EAF) slag (368 ± 10); ultrabasic mine resulted in a greater CO2 uptake and carbonate conversion compared to
tailings (40-250); and red mud (47 ± 8). Steelmaking slags thus have those of AOD slag, at essentially all processing conditions [33]. In our
substantial potential for CO2 emissions reduction via mineral carbo- subsequent work, seven high-purity minerals (åkermanite, bredigite,
nation of waste materials, and much research on this topic has been cuspidine, merwinite, srebrodolskite (Ca2Fe2O5), β- and γ-dicalcium
done in recent years [9,10]. silicate (C2S)) that are commonly found in ironmaking and steelmaking
The annual steel production exceeds 1.8 billion tonnes worldwide slags were synthesized and underwent carbonation experiments to un-
[11]. For every tonne of steel produced, 1.9 tonnes of CO2 is generated, derstand the affinity of the individual mineral phases of the slags to-
and as a result, the steel industry is one of the biggest sources of in- wards mineral carbonation [34]. The synthetic minerals were produced
dustrial CO2, accounting for about 7% of the total anthropogenic at high purity (>70 wt% target mineral) with small composition of
emissions to the atmosphere [12]. The reduction of these emissions for residual free oxides (<10 wt%). Even though we demonstrated that
climate change mitigation is difficult because the industry is heavily bredigite is the most favorable mineral phase for mineral carbonation,
dependent on either coke or coal for power generation. Since modern not all results of the synthetic minerals corresponded to those of the
integrated LD/converter steel plants produce about 90–100 kg of slag slags. For example, the carbonation conversion of the two C2S poly-
per tonne of steel, vast amounts of steelmaking slag are being co-pro- morphs was higher in CC slag than that of these synthetic minerals
duced [11]. Estimations of the world’s annual steelmaking slag pro- when both samples were carbonated via thin-film carbonation [34].
duction vary between 190 and 290 Mt [13]. High calcium oxide content The inconsistency of the results indicates that it is not only the mi-
of steelmaking slags makes them suitable for use as raw material in neralogical susceptibility towards mineral carbonation that affects the
mineral carbonation. With a calcium oxide content of 40–52% [14], carbonation reactivity of the slags, but also the particle morphology
ideally, this amount could store up to 170 Mt of CO2 per year by including the intra-particle location of the mineral phase components

2
Y.E. Chai, Q.R.S. Miller, H.T. Schaef et al. The Journal of Supercritical Fluids 171 (2021) 105191

[34]; solid solutions, degree of crystallinity and crystal defects, and The in situ XRD experimental apparatus and procedures are described
cooling hystory may be other important factors. For both studies, ex situ in detail elsewhere [35,36]. Briefly, the in situ XRD reactor was housed
X-ray powder diffraction (XRD) was performed on the carbonated in a Bruker D8 Discover X-ray powder diffractometer equipped with a
samples to analyze their mineral phases. As the analysis was conducted HI-STAR GADDS detector, an experimental configuration capable of
solely on the final products, it was not possible to determine reaction analysing mineral assemblages at elevated pressure and temperature
pathways, including identification of intermediate metastable phases or under a 45 kV and 200 mA Cu Kα-sourced microfocus X-ray beam. The
quantifying the rate of carbonation conversion of individual silicate diffractometer automatically scanned the sample for 200 s with an
phases. approximate nine second delay in between each 10–46°2θ pattern.
Herein, we report in situ XRD results for a series of carbonation The experiments were performed independently using two different
experiments conducted in a custom-built beryllium-capped reactor water-saturated (wet) CO2 phases: subcritical and supercritical. The
filled with either water-saturated (wet) subcritical (70 °C, 6 bar) CO2 equilibrium concentrations of H2O were 5.36 and 0.33 mol% for gas-
(CO2(g)), or wet supercritical (50 °C, 82 bar) CO2 (scCO2), and con- eous and supercritical conditions, respectively [37]. For experiments in
taining AOD slag, CC slag, or one of their six constituent synthetic which wet CO2(g) was supplied, temperature of 70 °C, CO2 pressure of
minerals: åkermanite, bredigite, cuspidine, merwinite, β-C2S, and 6 bar, and 40 μl of water was used, while 10 μl of water was placed in
γ-C2S. The carbonation products, which include various forms of cal- the reactor prior to pressurization at 50 °C with scCO2 at pressure of
cium and magnesium carbonates, were characterized both with respect 82 bar. Table S2 presents the reaction conditions used for carbonation
to their amounts and rate of formation and to the process conditions of each material, and nomenclature used to refer to each set of condi-
that promote their formation. The in situ XRD approach allowed to tions (EXP #1 to EXP #11). The nominal amount of mineral used in
determine reaction pathways of individual mineral phases and their each experiment was ~ 10 mg. A predetermined amount of water (40 or
transformation during carbonation, which subsequently led to identi- 10 μl) was deposited in the base of the reactor to ensure that the CO2
fication of the phases most susceptible to carbonation reaction. The link was always fully saturated with H2O [37]. A Model 100D ISCO syringe
between mineral composition, mineral structure and carbonation extent pump was used to adjust the pressure in the reactor and the tempera-
is also discussed to clarify the mineralogical influence on carbonation. ture of the reactor was maintained within ±1 °C for all experiments.
By knowing which mineral phases in stainless steel slags are susceptible The XRD patterns were analyzed using both GADDS XRD software
to reacting under certain conditions during mineral carbonation, (Bruker-AXS) and JADE XRD software (MDI). Each mineral phase was
steelmakers can use this information to tune the chemical composition identified based on the database provided by Mineralogy Database
of the slag and control the slag cooling path so that the cooled material (webmineral.com). Analysis of GADDS detector images provided in-
has a more desirable mineralogy for mineral carbonation. Also, by formation on crystallite microstructure [38] of single crystal or pow-
conducting in situ XRD analysis, the behaviour of each mineral during dered solid samples, such as phase identification, grain size, poly-
carbonation and the mineral phases that appear and disappear can be crystallinity, and texture [39].
been identified, which contributes to increasing confidence in the To provide quantitative insight into carbonate abundance and com-
predictability of carbonation effects and encourages future research for position, reacted subsamples recovered from in situ experiments were
a more feasible CO2 sequestration process at industrial scale. subjected to thermogravimetric analysis coupled with mass spectrometry
(TGA-MS). As described previously [40], sample weight changes during
2. Materials and methods thermal decomposition in an N2 atmosphere were measured with a
precision microbalance, while associated CO2 and H2O releases were
2.1. Synthetic minerals monitored by observing ion currents (m/z) 44 and 18, respectively. TGA-
MS analyses were performed with a Netzsch Instruments TG-209F1
The six target minerals were synthesized based on solid-state sin- thermogravimetric analyzer connected to a QMS 403 C Aëolos quadru-
tering methodology and conditions (sintering temperatures, duration, pole mass spectrometer. Each sample was heated in an alumina crucible
cooling path) found in published literature [34], with the goal of at 2 °C/min to 900 °C under 20 ml/min N2 flow. Eq. (1) was used to
achieving >70 wt% purity of the target mineral and <10 wt% free calculate the maximal CO2 uptake of each material, which uses Ca and
oxides. Table S1 shows the complete mineralogical composition of the Mg contents of the materials. In Eq. (1), QCO2 is the CO2 uptake de-
synthesized minerals. termined by TGA, expressed as grams CO2 captured per gram of initial
material mass; MW are elemental molar masses, and w are elemental
2.2. Stainless steelmaking slags mass fractions in the original material. Table 1 summarizes the Ca and
Mg mass fractions in the original materials and maximal CO2 uptake
The AOD and CC stainless steelmaking slags were obtained from the calculated.
production process of a Belgian steelworks, and had <500 µm particle
QCO 2
size. The slags were composed of high concentrations of calcium and MWCO2
moderate to low concentrations of magnesium. These concentrations CO2 uptake (% maximal) = wCa wMg
+ MW
contribute the alkaline properties of the slags and the reactivity towards MWCa Mg (1)
mineral carbonation [33,34]. The mineralogical compositions of the
slags were previously determined by quantitative X-ray diffraction
(QXRD). The main mineral phase of both AOD and CC slags is γ-C2S, with 3. Results
compositions of 24.1 and 38.6 wt%, respectively. Other mineral phases
identified include bredigite, cuspidine, β-C2S, merwinite, and periclase. The main research questions that have been investigated herein
Table S1 summarizes complete mineralogical composition of AOD and CC include: (i) do different silicates minerals that make up steelmaking
slags. Given their small but significant free lime content, prior to experi- slags have different rates of conversion under super/subcritical condi-
ments, the synthetic C2S samples were refreshed in an oven at 900 °C to tions; (ii) how do different silicates govern the type of carbonates that
remove initial trace carbonates that can form upon atmospheric exposure. form; (iii) do carbonates crystallize throughout the reaction or only
upon depressurization (i.e. pressure swing), and does this depend on
2.3. Pressurized in situ X-ray diffraction carbonation experiments whether super/subcritical CO2 is used; and (iv) do individual silicate
phases that make-up steelmaking slags behave differently when they
Time-resolved XRD experiments were conducted in a pressure- carbonate as pure mineral phases. To investigate the latter, the six
compatible static reactor with a beryllium cap and stainless-steel base. silicate minerals synthesized were compared against the two types of

3
Y.E. Chai, Q.R.S. Miller, H.T. Schaef et al. The Journal of Supercritical Fluids 171 (2021) 105191

Table 1
Ca and Mg mass fractions and maximal CO2.

Material CO2 Ca (wt%) Mg (wt%) Carbonate content, maximum (wt%) Carbonate content, TGA-MS (wt%) % maximal

β-C2S sc 46.55 0 51.1 14.6 30


γ-C2S sc 46.55 0 51.1 24.8 48
Cuspidine g 43.75 0 48.0 27.4 51
Bredigite g 41.67 3.60 52.3 18.9 37
Bredigite sc 41.67 3.60 52.3 24.5 48
AOD slag g 40.62 4.52 52.8 9.8 19
CC slag g 37.19 5.96 51.6 21.8 42

slags (AOD and CC). The carbonation reaction pathways involving Ca- Also, vaterite intensities (24.85, 27, and 32.78°2θ) began to increase in
rich minerals, Ca- and Mg-rich minerals, and stainless steelmaking slags the beginning of the reaction time. However, intensities at 24.85 and
are presented in Sections 3.1, 3.2, and 3.3, respectively. 27°2θ started to decrease in the middle of the reaction time. This de-
crease may have resulted from intensities of the wollastonite peaks
(25.28 and 26.88°2θ) decreasing in the middle of carbonation, hin-
3.1. Ca-rich minerals
dering the continuous increase in vaterite formation (24.85, 27, and
32.78°2θ), which then converted to more stable calcite. When cuspidine
The C2S minerals and cuspidine contain Ca as their alkaline earth
was carbonated using thin-film and slurry carbonation methods, for-
metal. According to the XRD diffractogram of β-C2S carbonation under
mation of vaterite or aragonite carbonates was not identified by XRD
wet CO2(g) conditions (EXP #1), β-C2S intensities (32.17 and 32.61°2θ)
[34]. The observed formation of vaterite and aragonite during cuspi-
immediately decreased within 1 h of reaction, and no further change in
dine carbonation may have been promoted by the in situ XRD method
intensity was observed (Fig. S1). At the same time, spurrite
used in this study, allowing detection of transient phases. The thermal
(Ca5(SiO4)2CO3) intensities (33.14, 33.57, and 33.85°2θ) dropped and
behavior according to the TGS-MS results was consistent with crystal-
did not change until the end of the reaction time. In terms of carbonates
line carbonates (Fig. S9) [41]. According to the calculated values
formation, calcite (29.4 and 39.4°2θ) and aragonite (26.22°2θ) intensities
summarized in Table 1, cuspidine achieved comparable CO2 uptake in
increased in the middle of the reaction time. When β-C2S was carbonated
wet CO2(g) as γ-C2S in wet scCO2 carbonation, while β-C2S performed
under wet scCO2 condition (EXP #2), the intensities of β-C2S and spurrite
poorer, indicating that the synthetic cuspidine is most susceptible to
decreased in the same manner as seen in EXP #1 (Fig. S2). However,
carbonation among the Ca-rich minerals tested. The GADDS images of
formation of aragonite was not detected, and sole formation of calcite
carbonated cuspidine indicated distinct calcite and aragonite forma-
indicates that carbonation in wet scCO2 was more thermodynamically
tion, as seen in Fig. 2. Fig. 2a is the GADDS image of unreacted syn-
stable than that in wet CO2(g). The TGA-MS results indicate its thermal
thetic cuspidine. The uniform Debye powder rings at the cuspidine
behavior regarding calcination was most consistent with that of calcite
peaks (27.2 and 29.1°2θ) indicate that cuspidine is a fine-grained
(Fig. S3) [41]. The GADDS images generated for EXP #2 showed a dis-
polycrystalline phase with a random orientation [39]. If the GADDS
tinct indication of calcite formation and absence of aragonite, being in
image exhibited “graininess,” consisting of discontinuous Debye
line with the other analysis results (Fig. S4).
powder rings, then it is consistent with the presence of relatively few
When γ-C2S was carbonated with wet CO2(g) (EXP #3) the in-
crystallites [39]. Fig. 2b is GADDS image after carbonation still under
tensities of the main peaks of γ-C2S, observed at 29.7 and 32.5°2θ in the
pressure, and Fig. 2c is after carbonation, depressurized. When pressure
QXRD, started to decrease immediately and quickly within 1 h of re-
was removed, the bright diffraction rings became more noticeable.
action, as well as the peaks of β-C2S (Fig. S5). In contrast, the intensities
Fig. 2c reveals that calcite and aragonite formed by carbonation, which
of the calcite peaks (29.4 and 39.4°2θ) increased gradually throughout
is consistent with the XRD results. Also, the vaterite peaks are absent in
the reaction time. Interestingly, formation of aragonite, which was
these figures as explained by their disappearance observed in Fig. 1c in
formed during β-C2S carbonation in wet CO2(g) (EXP #1), was not de-
the middle of the carbonation reaction.
tected. The carbonation of γ-C2S in wet scCO2 (EXP #4) showed similar
decrease in γ-C2S and β-C2S intensities, however, carbonate formation
was not seen (Fig. S6). According to [34], slower carbonation kinetics is 3.2. Ca- and Mg-rich minerals
suggested for C2S polymorphs as they benefit more from extended
carbonation time, and the small CO2 uptakes of these materials under Bredigite, merwinte, and åkermanite contain both Ca and Mg in
thin-film carbonation are caused by their free oxides contents. Con- their elemental composition. During bredigite carbonation in wet
tradicting the results of XRD, TGA-MS result of γ-C2S carbonated under CO2(g) (EXP #6), intensities of bredigite (32.78 and 33.63°2θ) and
wet scCO2 condition (EXP #4) suggests decomposition of calcite merwinite (33.32, 33.52, and 33.76°2θ) peaks decreased immediately
(Fig. S7) [41], and GADDS images generated for EXP #4, at the end of in the beginning and continued to gradually drop throughout the re-
the experiment after depressurization, showed the increased presence action time (Fig. S10). Huntite (31.55°2θ) and åkermanite (31.24°2θ)
of calcite (Fig. S8); thus both results imply formation of calcite. It is intensities decreased moderately throughout the reaction time. Arago-
justifiable to believe that calcite formation did occur during carbona- nite intensities (26.22 and 27.22°2θ) increased gradually throughout
tion of γ-C2S in scCO2, but the increased intensities of the calcite peaks the reaction time and that of hydromagnesite (30.82°2θ) rose at a
were too minimal to be detected in the XRD diffractograms, or the slower rate compared to aragonite. The dominant formation of arago-
calcite precipitated/crystallized only upon depressurization. nite was expected as similar results were obtained by bredigite carbo-
Cuspidine carbonation in wet CO2(g) (EXP #5) showed simulta- nation using thin-film and slurry carbonation methods, confirmed by
neously occurring transformations throughout the reaction time, as Bodor et al. [34]. As seen in TGA-MS results, the early decomposition
seen in Fig. 1. Fig. 1a presents the 3D XRD diffractogram and Fig. 1b is (~320–420 °C) may have resulted from thermal behaviour of amor-
the 2D waterfall plot of Fig. 1a. The intensities of the cuspidine peaks phous hydromagnesite (Fig. S11). During carbonation of bredigite in
(27.24 and 29.14°2θ) immediately decreased in the beginning of the wet scCO2 (EXP #7), similar disappearance in bredigite and merwinite
reaction time and continued to gradually drop, while calcite (29.4 and and growth in aragonite and hydromagnesite intensities were seen
39.4°2θ) and aragonite (26.22°2θ) gradually formed throughout the (Fig. S12). The TGA-MS results are very similar to those of EXP #6
reaction time, and the sample became enriched in fluorite (28.28°2θ). (Fig. S13). The GADDS images for bredigite carbonated in wet CO2(g)

4
Y.E. Chai, Q.R.S. Miller, H.T. Schaef et al. The Journal of Supercritical Fluids 171 (2021) 105191

Fig. 1. 3D (a) and 2D waterfall (b) plots for time-resolved diffractograms of synthetic cuspidine carbonation in water-saturated CO2(g) at 70 °C and 6 bar.

(EXP #6) exhibited uniform Debye lines at aragonite peak (26.2°2θ) but the synthetic minerals tested in this study, while bredigite was found to
not hydromagnesite (Fig. S14). Similar GADDS images were produced be the most reactive mineral [34].
for bredigite carbonated in wet scCO2 (EXP #7) as seen in Figs. 2d to 2f. During carbonation of synthetic åkermanite in wet CO2(g) (EXP #9),
When Figs. 2d and 2e are compared, distinct rings for aragonite ap- the intensities of åkermanite (28.87 and 31.14°2θ) and diopside (30.88
peared. In Fig. 2f, after depressurization, the aragonite rings became and 35.48°2θ) decreased in the beginning of the reaction, however,
more visible. The absence of hydromagnesite peaks in these figures formation of carbonates was not observed. This observation is reflected
could have resulted because its growth during carbonation was minimal on the XRD diffractograms of åkermanite carbonation, provided as
compared to other phases with greater intensities. Fig. S16. This result is in contrast with åkermanite carbonation using
During merwinite carbonation in wet CO2(g) (EXP #8) (Fig. S15), thin-film and slurry carbonation methods by Bodor et al. [34].
similar XRD results to bredigite carbonation were seen (EXP #6)
(Fig. S10). The intensities of merwinite (33.32, 33.52, and 33.76°2θ) 3.3. Stainless steelmaking slags
immediately decreased as the reaction started and continued to drop
gradually. The åkermanite intensities (31.24°2θ), a secondary phase in During carbonation of AOD slag in wet CO2(g) (EXP #10), the in-
the synthetic merwinite, sharply decreased in the beginning of the re- tensities of γ-C2S (29.7 and 32.5°2θ), cuspidine (27.34 and 29.14°2θ),
action. Aragonite (26.22 and 27.22°2θ) formed gradually throughout bredigite (32.78 and 33.63°2θ), and merwinite (33.32, 33.52, and
the reaction time and hydromagnesite (30.82°2θ) started to appear in 33.76°2θ) responded immediately (Fig. S17). Such mineral phases were
the middle of the reaction time. The formation of aragonite and hy- the main component of fresh AOD slags according to the QXRD results.
dromagnesite in bredigite and merwinite carbonation was encouraged The dominant conversions of bredigite and γ-C2S during AOD carbo-
by the presence of Mg in their original composition. The presence of Mg nation were also seen in thin-film carbonation by Santos et al. [33].
is known to inhibit precipitation of calcium carbonate as the more Near the end of the reaction time, monohydrocalcite (20.49°2θ) formed
stable calcite and to promote aragonite formation [42]. According to and formation of calcite or aragonite was not exhibited by the XRD
our earlier study on merwinite carbonation through thin-film and slurry diffractograms. Monohydrocalcite is an intermediate product in the
carbonation, merwinite had the slowest carbonation conversion out of transition of amorphous calcium carbonate (ACC) to stable calcite or

5
Y.E. Chai, Q.R.S. Miller, H.T. Schaef et al. The Journal of Supercritical Fluids 171 (2021) 105191

Fig. 2. GADDS images of: (a) unreacted cuspidine; (b) carbonated cuspidine in wet CO2(g) at 70 °C and 6 bar with 40 μl of H2O before depressurization and (c) after
depressurization; (d) unreacted bredigite, (e) carbonated bredigite in wet scCO2 at 50 °C and 82 bar with 10 μl of H2O before depressurization and (f) after de-
pressurization.

aragonite. The TGA-MS results showed that mass drop of 8.8% between spots, which was not observed in the synthetic materials, in both un-
300 and 620 °C (Fig. 3a). Calcination began at 300 °C and continued reacted and carbonated slags (Figs. S19 and S20). These features in-
gradually until 426 °C. Early decomposition (~330–420 °C) depicts dicate the presence of relatively few crystallites and coarse-grained
calcination of ACC [43,44]. In coincidence, between such temperature crystallites (>40 µm) [39]. The GADDS images generated for EXP #10
range, significant release of H2O was detected. The early calcination and EXP #11 did not show distinct indication of carbonate formation,
and simultaneous release of H2O is consistent with the behavior of however, the peaks of carbonates may have been obscured by nearby
hydrated ACC. Then, the mass drop ceased until 560 °C and then ap- peaks. When the results from TGA-MS are compared, CC slag resulted in
peared a second peak, which was much steeper, at 622 °C with a sharp a greater CO2 uptake compared to AOD slag (Table 1), in agreement
decrease. This behavior is consistent with monohydrocalcite [45], with thin-film and slurry carbonation experiments done previously
which was also proven to emerge in thin-film carbonation of AOD slag [33]. Cuspidine and bredigite exist in higher amounts in fresh AOD slag
by Santos et al. [33]. than fresh CC slag, while CC slag contains more C2S polymorphs in its
Carbonation of CC slag in wet CO2(g) (EXP #11) showed decrease in composition, as the values can be found in Table S1. Despite compar-
β-C2S (32.17 and 32. 61°2θ), γ-C2S (29.7 and 32.5°2θ), bredigite (32.78 able CO2 uptake of individual minerals (Table 1), it appears that cus-
and 33.63°2θ), and merwinite (33.32, 33.52, and 33.76°2θ) intensities pidine and bredigite are poorly reactive in AOD slag carbonation, while
(Fig. S18). The dominance of C2S polymorphs in carbonation conver- C2S largely contribute to the higher CO2 uptake achieved by CC slag.
sion is in parallel with the results found by Santos et al. [33], in which
CC slag was carbonated by thin-film carbonation method. Near the end 4. Discussion
of the reaction, calcite intensities (29.4°2θ) increased. However, neither
Ca-Mg-carbonates nor Mg-carbonates formation were detected despite In Section 4.1, supercritical and subcritical CO2 carbonation are
the decrease in bredigite and merwinite peaks. The reacted Mg could contrasted, and in Section 4.2, links between the crystal structure of
have been incorporated in the calcite lattice, widening the calcite peak silicate minerals and their behavior during mineral carbonation are
[46]. Similar to the AOD slags, CC slag experienced two peaks of CO2 discussed.
release according to the TGA-MS results shown in Fig. 3b. The first and
minor one occurred at 445 °C and the second one was at 620 °C. 4.1. Subcritical vs. supercritical CO2
However, distinct H2O release was not observed. This phenomenon may
imply that the AOD slag was more hydrated than CC slag, but both slags Reacting minerals with scCO2 has been gaining attention to evaluate
contained ACC and crystalline carbonates formed after carbonation in the potential impacts it has to reservoir and caprock systems [47–50].
wet CO2(g). Studying carbonation of forsterite in wet scCO2 determined that thin
The GADDS images of the slags (AOD and CC) presented the liquid water film formation on the mineral surface resulting from water-
“graininess” in the diffraction rings with random bright diffraction bearing scCO2 is crucial to the carbonation rate and extent [51]. The

6
Y.E. Chai, Q.R.S. Miller, H.T. Schaef et al. The Journal of Supercritical Fluids 171 (2021) 105191

Fig. 3. TGA-MS curves of: (a) carbonated AOD slag; (b) carbonated CC slag in wet CO2(g) at 70 °C and 6 bar with 40 μl of H2O.

investigations discovered the impact of dissolved water concentration produced due to difference in reaction conditions, which could have
in scCO2 on carbonation performance and that increased water content increased passivation by silica layers, limiting carbonation reaction in
increases reactivity. In the present study, carbonation in wet scCO2 with wet CO2(g), as commented by Daval et al. [52]. It is too early to make a
constant concentration of water was conducted for β-C2S, γ-C2S, and conclusion on the effect of using scCO2 as opposed to CO2(g) as only a
bredigite carbonation to investigate the reaction of such materials in few experimental results were produced from scCO2 carbonation, but it
scCO2 and study how it affects the reaction compared to carbonation in is confirmed that steel slag mineral phases carbonate in scCO2, thus
wet CO2(g). there is potential for the integration of mineral carbonation processes
Overall, contradicting effects were seen in C2S minerals. During β- with processes that generate scCO2 as a byproduct, such as supercritical
C2S carbonation in wet CO2(g) (EXP #1), calcite formed as well as water oxidation and gasification [62].
aragonite (Fig. S1). When it was exposed to wet scCO2 (EXP #2), for-
mation of only calcite was detected, indicating that carbonation was 4.2. Crystal structure and mineral carbonation
more thermodynamically stable than that in wet CO2(g) (Fig. S2).
During carbonation of γ-C2S in both wet CO2(g) and wet scCO2 (EXP Crystal structure of minerals affects the carbonation kinetics and
#4), it was verified that calcite formed (Fig. S6). In bredigite carbo- extent of carbonation. The crystal structure of the unreacted minerals is
nation, significant difference between using wet CO2(g) and using wet responsible for their dissolution in H2O [53], and it influences the
scCO2 during for carbonation was not seen as formation of aragonite breaking of ionic bonds to release the alkaline earth metal ions. The
and hydromagnesite was exhibited in both conditions (Figs. S10 and minerals that were subject to this study are classified as two silicate
S12). According to Table 1, bredigite carbonation in wet scCO2 types: nesosilicate and sorosilicate. The C2S minerals, bredigite, and
achieved greater CO2 uptake than in wet CO2(g). This result was merwinite belong to nesosilicate group, and cuspidine and åkermanite

7
Y.E. Chai, Q.R.S. Miller, H.T. Schaef et al. The Journal of Supercritical Fluids 171 (2021) 105191

are in the sorosilicate group. Nesosilicates constitute an isolated silicon chemistry with fluxes and by control of slag cooling path.
tetrahedra, in the form of orthosilicate ion (SiO44-), which is connected The set of experimental conditions and approaches used in this work
only by interstitial cations [54]. The dissolution of such silicates occurs is unique in relation to the present literature, and the findings help this
in multiple steps where Ca‒O and Mg‒O bonds break followed by the research field progress towards better understanding of the geochem-
release of SiO4 tetrahedra or double tetrahedra, without the need to istry and mineralogy that govern the carbonation of complex calcium-
break Si‒O bonds [55]. Sorosilicates contain a bridging oxygen and magnesium-silicates mixtures. The in situ XRD technique clearly
(Si‒O‒Si) resulting in double tetrahedra, in the form of isolated pyr- demonstrated the minerals that are susceptible to mineral carbonation
osilicate anion (Si2O76-). The strong Si‒O‒Si bonds make these minerals in steelmaking slags, the transient phases that form and reactions that
less susceptible to dissolution and chemical reaction [56]. cease early or commence late, and the relative conversion rates and
In terms of the minerals synthesized in this study, the C2S minerals degrees of reaction conversion, which is an information that cannot be
were expected to have the simplest dissolution step as the only bond obtained in the conventional ex situ XRD approach. It also demon-
required to be broken is the Ca‒O bond to liberate the SiO4 tetrahedra strated how supercritical and subcritical CO2, in absence of liquid
[55]. In C2S, all Si atoms are part of a silicate group and all Ca atoms are aqueous phase, promote conversion of silicates into carbonates. Mineral
ionic in character [57]. The weaker Ca‒O ionic bond of β-C2S compared carbonation without liquid aqueous phase is attractive since it facil-
to γ-C2S contributes to its elevated reactivity and leads to higher car- itates recovery of the exothermic heat of carbonation reaction, which
bonation extent. β-C2S has two classes of Ca atom in which one class is can be interrupted by the large heat capacity of the water phase in
six-coordinated and the second is eight-coordinated [58]. The Ca atoms aqueous systems. Developing better mechanistic understanding of in-
in γ-C2S are six-coordinated and the Ca‒O bond of the Ca in the poly- dividual mineral phases on the reaction kinetics will help make a better
hedral coordination sites is less stable and breaks down before that in use of the steelmaking slags as the feedstock for mineral carbonation
the octahedra sites, in which the presence of distortions that disturb the and make predictive models for utilization of the steelmaking slags for
structure of a regular octahedron increases its stability [58]. The simple CO2 sequestration.
extraction of the Ca present in C2S contributes to quick conversion of
C2S, as seen during XRD carbonation (Figs. S1, S2, S5 and S6). Declaration of Competing Interest
Bredigite and merwinite are also nesosilicates but they contain both
Ca and Mg, meaning Mg‒O bond exists in their crystal structure as The authors declare that they have no known competing financial
well [63,64]. In contrast to C2S carbonation behavior, both bredigite interests or personal relationships that could have appeared to influ-
and merwinite showed continuous conversion throughout the reaction. ence the work reported in this paper.
For example, in the XRD results during bredigite carbonation in wet
CO2(g) (EXP #6) both bredigite and merwinite intensities decreased Acknowledgments
immediately in the beginning of the reaction and the gradual drop
continued until the end of the reaction time (Figs. S10 and S15). It may RMS would like to acknowledge the Discovery Grant research
be predicted that the Mg, which has smaller ion size compared to Ca, funding provided by the Natural Sciences and Engineering Research
dissolved slower in water leading to gradual carbonation rate [59]. Council of Canada, Canada (NSERC), and the travel funding provided
Cuspidine and åkermanite are sorosilicates and have double Si tet- by the Fonds Wetenschappelijk Onderzoek, Belgium (FWO). QM and
rahedra that share an oxygen atom (i.e. pyrosilicate anion), but cuspi- HTS acknowledge support from the U.S. Department of Energy, USA
dine also contains F ions in its structure, forming ionic bond with (US DOE), Office of Science, Office of Basic Energy Sciences (BES),
Ca [64,65]. During XRD carbonation, cuspidine intensity immediately Chemical Sciences, Geosciences, and Biosciences Division through its
decreased and continued to gradually drop until the end of the reaction Geosciences program at Pacific Northwest National Laboratory (PNNL),
(Fig. 1). It can be suggested that Ca loosely bound to O led to immediate Richland, Washington, and the US DOE Office of Fossil Energy at PNNL
carbonation, while the breaking of Ca–O‒Si bonds and release of Ca through the National Energy Technology Laboratory, Morgantown,
bonded to F occurred subsequently, thus contributing to a more gradual West Virginia.
rate of carbonation and fluorite enrichment.
The structure of åkermanite consists of (MgSi2O7)4- layer of inter- Appendix A. Supporting information
connected tetrahedra and the Ca atoms lie between these layers [60].
Mg has a 4-fold tetrahedral coordination and Ca an 8-fold polyhedral Supplementary data associated with this article can be found in the
coordination [61]. During XRD carbonation, åkermanite did not show online version at doi:10.1016/j.supflu.2021.105191.
any transformation except for the decrease in åkermanite and diopside
peaks in the very beginning of the reaction time nor any formation of References
carbonation was observed (Fig. S16). As carbonation of åkermanite was [1] IPCC, Climate Change 2007: Synthesis Report. Contribution of Working Groups I, II
seen in previous studies [34], it was the reaction conditions that did not and III to the Fourth Assessment Report of the Intergovernmental Panel on Climate
favor carbonation of this specific mineral rather than its crystal struc- Change [Core Writing Team, Pachauri, R.K and Reisinger, A. (eds.)]. IPCC,Geneva,
Switzerland, 2007, pp. 104.
ture. In our previous work, in which the same synthetic minerals were [2] W.J.J. Huijgen, R.N.J. Comans, Mineral CO2 sequestration by steel slag carbona-
carbonated through thin-film and slurry carbonation, åkermanite tion, Environ. Sci. Technol. 39 (24) (2005) 9676–9682, https://doi.org/10.1021/
showed moderate carbonation conversion and CO2 uptake compared to es050795f
[3] K.S. Lackner, A guide to CO2 sequestration, Science 300 (5626) (2003) 1677–1678,
others. https://doi.org/10.1126/science.1079033
[4] H.-J. Ho, A. Iizuka, E. Shibata, H. Tomita, K. Takano, T. Endo, CO2 utilization via
5. Conclusions direct aqueous carbonation of synthesized concrete fines under atmospheric pres-
sure, ACS Omega 5 (2020) 15877–15890, https://doi.org/10.1021/acsomega.
0c00985
Through identifying the behavior of individual minerals in the [5] E.J. Moon, Y.C. Choi, Carbon dioxide fixation via accelerated carbonation of ce-
steelmaking slags during carbonation, new insights into utilizing ment-based materials: potential for construction materials applications, Constr.
Build. Mater. 199 (2019) 676–687, https://doi.org/10.1016/j.conbuildmat.2018.
steelmaking slags as a mean for mineral carbonation were observed.
12.078
The motivation of this work was to use slags as carbon sink, which can [6] A. Di Maria, M. Salman, M. Dubois, K. Van Acker, Life cycle assessment to evaluate
benefit steelmakers by reducing their emissions and also by stabilizing the environmental performance of new construction material from stainless steel
their slags. In addition, by knowing which silicate minerals that make- slag, Int. J. Life Cycle Assess. 23 (2018) 2091–2109, https://doi.org/10.1007/
s11367-018-1440-1
up slags are most reactive to carbonation, this work informs steel- [7] F. Bourgeois, P. Laniesse, M. Cyr, C. Julcour, Definition and exploration of the in-
makers of what slag composition to aim for when tuning the slag tegrated CO2 mineralization technological cycle, Front. Energy Res. 8 (2020) 113,

8
Y.E. Chai, Q.R.S. Miller, H.T. Schaef et al. The Journal of Supercritical Fluids 171 (2021) 105191

https://doi.org/10.3389/fenrg.2020.00113 and chemical assessment, Eur. J. Miner. 15 (2013) 533–549, https://doi.org/10.


[8] P. Renforth, The negative emission potential of alkaline materials, Nat. Commun. 1127/0935-1221/2013/0025-2300
10 (2019) 1401, https://doi.org/10.1038/s41467-019-09475-5 [35] M.A. Sinnwell, Q.R.S. Miller, L. Liu, J. Tao, M.E. Bowden, L. Kovarik, D. Barpaga,
[9] X. Zhang, J. Chen, J.J. Jiang, J. Li, R.D. Tyagi, R.Y. Surampalli, The potential utilization Y. Han, R. Kishan Motkuri, M.L. Sushko, H.T. Schaef, P.K. Thallapally, Kinetics and
of slag generated from iron- and steelmaking industries: a review, Environ. Geochem. mechanisms of ZnO to ZIF‐8 transformations in supercritical CO2 revealed by in situ
Health 42 (2020) 1321–1334, https://doi.org/10.1007/s10653-019-00419-y X‐ray diffraction, ChemSusChem 13 (10) (2020) 2602–2612, https://doi.org/10.
[10] G. Kashwani, A review of CO2 sequestration via fly ash and steel slag, Emir. Eng. 1002/cssc.202000434
Abroad Lett. 1 (2020) 1–11. [36] H.T. Schaef, C.F. Windisch Jr., B.P. McGrail, P.F. Martin, K.M. Rosso, Brucite
[11] World Steel Association, Steel statistical yearbook 2019: concise version, Worldsteel [Mg(OH2)] carbonation in wet supercritical CO2: an in situ high pressure X-ray
Committee on Economic Studies, Brussels, 2019. https://www.worldsteel.org/steel- diffraction study, Geochim. Cosmochim. Acta 75 (2011) 7458–7471, https://doi.
by-topic/statistics/steel-statistical-yearbook.html. (Accessed 18 July 2020). org/10.1016/j.gca.2011.09.029
[12] Y. Kim, E. Worrell, International comparison of CO2 emission trends in the iron and [37] N. Spycher, K. Pruess, J. Ennis-King, CO2–H2O mixtures in the geological seques-
steel industry, Energy Policy 30 (10) (2002) 827–838, https://doi.org/10.1016/ tration of CO2. I. Assessment and calculation of mutual solubilities from 12 to 100
S0301-4215(01)00130-6 °C and up to 600 bar, Geochim. Cosmochim. Acta 67 (2003) 3015–3031, https://
[13] United States Geological Surveys (USGS), Mineral commodity summaries 2019, doi.org/10.1016/S0016-7037(03)00273-4
USGS and United States Department of Interior, 2019. https://www.usgs.gov/ [38] B. Seuss, J. Titschack, S. Seifert, J. Neubauer, A. Nützel, Oxygen and stable carbon
centers/nmic/mineral-commodity-summaries. (Accessed 18 July 2020). isotopes from a nautiloid from the middle Pennsylvanian (Late Carboniferous)
[14] U.S. Department of Transportation, Federal Highway Administrations, User guide- impregnation Lagerstätte ‘Buckhorn Asphalt Quarry’ — Primary paleo-environ-
lines for waste and by product materials in pavement construction, 1997. https:// mental signals versus diagenesis, Palaeogeogr. Palaeoecol. 319–320 (2012) 1–15,
www.fhwa.dot.gov/publications/research/infrastructure/pavements/97148/ https://doi.org/10.1016/j.palaeo.2011.12.008
index.cfm. (Accessed 18 July 2020). [39] R.L. Flemming, Micro X-ray diffraction (μXRD): a versatile technique for char-
[15] S. Eloneva, S. Teir, J. Salminen, C.J. Fogelholm, R. Zevenhoven, Fixation of CO2 by acterization of Earth and planetary materials, Can. J. Earth Sci. 44 (2007)
carbonating calcium derived from blast furnace slag, Energy 33 (9) (2008) 1333–1346, https://doi.org/10.1139/E07-020
1461–1467, https://doi.org/10.1016/j.energy.2008.05.003 [40] Q.R.S. Miller, J.P. Kaszuba, S.N. Kerisit, H.T. Schaef, M.E. Bowden, B.P. McGrail,
[16] B. Das, S. Prakash, P.S.R. Reddy, V.N. Misra, An overview of utilization of slag and K.M. Rosso, Emerging investigator series: ion diffusivities in nanoconfined inter-
sludge from steel industries, Resour. Conserv. Recycl. 50 (1) (2007) 40–57, https:// facial water films contribute to mineral carbonation thresholds, Environ. Sci. Nano
doi.org/10.1016/j.resconrec.2006.05.008 7 (4) (2020) 1068–1081, https://doi.org/10.1039/c9en01382b
[17] R.M. Santos, D. Ling, A. Sarvaramini, M. Guo, J. Elsen, F. Larachi, G. Beaudoin, [41] R. Frost, M. Hales, W. Martens, Thermogravimetric analysis of selected group (II)
B. Blanpain, T. Van Gerven, Stabilization of basic oxygen furnace slag by hot-stage carbonate minerals—implication for the geosequestration of greenhouse gases,
carbonation treatment, Chem. Eng. J. 203 (2012) 239–250, https://doi.org/10. J. Therm. Anal. Calorim. 95 (3) (2009) 999–1005, https://doi.org/10.1007/
1016/j.cej.2012.06.155 s10973-008-9196-7
[18] G.C. Ulubeyli, R. Artir, Sustainability for blast furnace slag: use of some construc- [42] R.A. Berner, The role of magnesium in the crystal growth of calcite and aragonite
tion wastes, Procedia Soc. Behav. Sci. 195 (2015) 2191–2198, https://doi.org/10. from sea water, Geochim. Cosmochim. Acta 39 (4) (1975) 489–504, https://doi.
1016/j.sbspro.2015.06.297 org/10.1016/0016-7037(75)90102-7
[19] P. Chaurand, J. Rose, V. Briois, L. Olivi, J.L. Hazemann, O. Proux, J. Domas, [43] N. Koga, Y. Yamane, Thermal behaviors of amorphous calcium carbonates prepared
J.-Y. Bottero, Environmental impacts of steel slag reused in road construction: a in aqueous and ethanol media, J. Therm. Anal. Calorim. 94 (2) (2008) 379–387,
crystallographic and molecular (XANES) approach, J. Hazard. Mater. 139 (3) https://doi.org/10.1007/s10973-008-9110-3
(2007) 537–542, https://doi.org/10.1016/j.jhazmat.2006.02.060 [44] Q.R.S. Miller, C.J. Thompson, J.S. Loring, C.F. Windisch, M.E. Bowden, D.W. Hoyt,
[20] C. Liu, S. Huang, B. Blanpain, M. Guo, Optimization of mineralogy and micro- J.Z. Hu, B.W. Arey, K.M. Rosso, H.T. Schaef, Insights into silicate carbonation
structure of solidified basic oxygen furnace slag through SiO2 addition or atmo- processes in water-bearing supercritical CO2 fluids, Int. J. Greenh. Gas. Control 15
sphere control during hot-stage slag treatment, Metall. Mater. Trans. B 50B (2019) (2013) 104–118, https://doi.org/10.1016/j.ijggc.2013.02.005
210–218, https://doi.org/10.1007/s11663-018-1444-z [45] M. Neuman, M. Epple, Monohydrocalcite and its relationship to hydrated amor-
[21] R. Baciocchi, G. Costa, A. Polettini, R. Pomi, Effects of thin-film accelerated car- phous calcium carbonate in biominerals, Eur. J. Inorg. Chem. 2007 (14) (2007)
bonation on steel slag leaching, J. Hazard. Mater. 286 (2015) 369–378, https://doi. 1953–1957, https://doi.org/10.1002/ejic.200601033
org/10.1016/j.jhazmat.2014.12.059 [46] M.A. Boone, P. Nielsen, T. Decock, M.N. Boone, M. Quaghebeur, V. Cnudde,
[22] B. Liu, J. Li, Z. Wang, Y. Zeng, Q. Ren, Long-term leaching characterization and Monitoring of stainless-steel slag carbonation using X-ray computed micro-
geochemical modeling of chromium released from AOD slag, Environ. Sci. Pollut. tomography, Environ. Sci. Technol. 48 (1) (2014) 674–680, https://doi.org/10.
Res. 27 (2020) 921–929, https://doi.org/10.1007/s11356-019-07008-7 1021/es402767q
[23] Y.J. Wang, Y.N. Zeng, J.G. Li, Y.Z. Zhang, Y.J. Zhang, Q.Z. Zhao, Carbonation of [47] O. Regnault, V. Lagneau, H. Catalette, H. Schneider, Experimental study of pure
argon oxygen decarburization stainless steel slag and its effect on chromium mineral phases/supercritical CO2 reactivity. Implications for geological CO2 se-
leachability, J. Clean. Prod. 256 (2020) 120377, https://doi.org/10.1016/j.jclepro. questration, CR Geosci. 337 (2005) 1331–1339, https://doi.org/10.1016/j.crte.
2020.120377 2005.07.012
[24] E.R. Bobicki, Q. Liu, Z. Xu, H. Zeng, Carbon capture and storage using alkaline [48] H. Lin, T. Fujii, R. Takisawa, T. Takahashi, T. Hashida, Experimental evaluation of
industrial wastes, Prog. Energy Combust. Sci. 38 (2012) 301–320, https://doi.org/ interactions in supercritical CO2/water/rock minerals system under geologic CO2
10.1016/j.pecs.2011.11.002 sequestration conditions, J. Mater. Sci. 43 (7) (2008) 2307–2315, https://doi.org/
[25] P. Librandi, G. Costa, S. Stendardo, R. Baciocchi, Carbonation of BOF slag in a 10.1007/s10853-007-2029-4
rotary kiln reactor in view of the scale-up of the wet route process, Environ. Prog. [49] E. Ilton, H. Schaef, O. Qafoku, K. Rosso, A. Felmy, In situ X-ray diffraction study of
Sustain. Energy 38 (3) (2019) e13140, https://doi.org/10.1002/ep.13140 Na+ saturated montmorillonite exposed to variably wet super critical CO2, Environ.
[26] P. Librandi, P. Nielsen, G. Costa, R. Snellings, M. Quaghebeur, R. Baciocchi, Sci. Technol. 46 (7) (2012) 4241–4248, https://doi.org/10.1021/es300234v
Mechanical and environmental properties of carbonated steel slag compacts as a [50] X. Wang, V. Alvarado, N. Swoboda-Colberg, J.P. Kaszuba, Reactivity of dolomite in
function of mineralogy and CO2 uptake, J. CO2 Util. 33 (2019) 201–214, https:// water-saturated supercritical carbon dioxide: significance for carbon capture and
doi.org/10.1016/j.jcou.2019.05.028 storage and for enhanced oil and gas recovery, Energ. Convers. Manag. 65 (2013)
[27] S.O. Omale, T.S.Y. Choong, L.C. Abdullah, S.I. Siajam, M.W. Yip, Utilization of 564–573, https://doi.org/10.1016/j.enconman.2012.07.024
Malaysia EAF slags for effective application in direct aqueous sequestration of [51] H.T. Schaef, B.P. McGrail, J.L. Loring, M.E. Bowden, B.W. Arey, K.M. Rosso,
carbon dioxide under ambient temperature, Heliyon 5 (2019) e02602, https://doi. Forsterite [Mg2SiO4)] carbonation in wet supercritical CO2: an in situ high-pressure
org/10.1016/j.heliyon.2019.e026 X-ray diffraction study, Environ. Sci. Technol. 47 (2013) 174–181, https://doi.org/
[28] H. Zhang, Y. Lu, J. Dong, L. Gan, Z. Tong, Roles of mineralogical phases in aqueous 10.1021/es301126f
carbonation of steelmaking slag, Metals 6 (2016) 117, https://doi.org/10.3390/ [52] D. Daval, O. Sissmann, N. Menguy, G.D. Saldi, F. Guyot, I. Martinez, J. Corvisier,
met6050117 B. Garcia, I. Machouk, K.G. Knauss, R. Hellmann, Influence of amorphous silica
[29] B. Park, E.-J. Moon, Y.C. Choi, Investigation of microstructure and mechanical layer formation on the dissolution rate of olivine at 90 °C and elevated pCO2, Chem.
performance of carbon-capture binder using AOD stainless steel slag, Constr. Build. Geol. 284 (1) (2011) 193–209, https://doi.org/10.1016/j.chemgeo.2011.02.021
Mater. 242 (2020) 118174, https://doi.org/10.1016/j.conbuildmat.2020.118174 [53] J.M. Valverde, P.E. Sanchez-Jimenez, L.A. Perez-Maqueda, M.A.S. Quintanilla,
[30] Z.X. Chen, S.H. Chu, Y.S. Lee, H.S. Lee, Coupling effect of γ-dicalcium silicate and J. Perez-Vaquero, Role of crystal structure on CO2 capture by limestone derived
slag on carbonation resistance of low carbon materials, J. Clean. Prod. 262 (2020) CaO subjected to carbonation/recarbonation/calcination cycles at Ca-looping con-
121385, https://doi.org/10.1016/j.jclepro.2020.121385 ditions, Appl. Energy 125 (2014) 254–275, https://doi.org/10.1016/j.apenergy.
[31] F. Engström, Mineralogical Influence on Leaching Behaviour of Steelmaking Slags, 2014.03.065
Doctoral Dissertation, Luleå University of Technology, 2010. [54] S. Lavikko, O. Eklund, The role of the silicate groups in the extraction of Mg with
[32] R.M. Santos, D. François, G. Mertens, J. Elsen, T. Van Gerven, Ultrasound-in- the ÅA route method, J. CO2 Util. 16 (2016) 466–473, https://doi.org/10.1016/j.
tensified mineral carbonation, Appl. Therm. Eng. 57 (1–2) (2013) 154–163, https:// jcou.2016.10.012
doi.org/10.1016/j.applthermaleng.2012.03.035 [55] R. Ragipani, S. Bhattacharya, A.K. Suresh, Towards efficient calcium
[33] R.M. Santos, J. Van Bouwel, E. Vandevelde, G. Mertens, J. Elsen, T. Van Gerven, extraction from steel slag and carbon dioxide utilisation via pressure-swing
Accelerated mineral carbonation of stainless steel slags for CO2 storage and waste mineral carbonation, React. Chem. Eng. 4 (2019) 52–66, https://doi.org/10.
valorization: Effect of process parameters on geochemical properties, Int. J. Greenh. 1039/c8re00167g
Gas. Control 17 (2013) 32–45, https://doi.org/10.1016/j.ijggc.2013.04.004 [56] M.H. Fasihnikoutalab, S. Pourakbar, R.J. Ball, B.K. Huat, The effect of olivine content
[34] M. Bodor, R.M. Santos, L. Kriskova, J. Elsen, M. Vlad, T. Van Gerven, Susceptibility and curing time on the strength of treated soil in presence of potassium hydroxide, Int.
of mineral phases of steel slags towards carbonation: mineralogical, morphological J. Geosynth. Ground Eng. 3 (2017) 1–10, https://doi.org/10.1007/s40891-017-0089-3

9
Y.E. Chai, Q.R.S. Miller, H.T. Schaef et al. The Journal of Supercritical Fluids 171 (2021) 105191

[57] E. Durgun, H. Manzano, R.J.M. Pellengq, J.C. Grossman, Understanding and con- surfaces, Acta Mater. 57 (2009) 5303–5313, https://doi.org/10.1016/j.actamat.
trolling the reactivity of the calcium silicate phases from first principles, Chem. 2009.07.023
Mater. 24 (2012) 1262–1267, https://doi.org/10.1021/cm203127m [62] S. Abdpour, R.M. Santos, Recent advances in heterogeneous catalysis for super-
[58] X. Feng, X. Min, C. Tao, Study on the structure and characteristic of dicalcium critical water oxidation/gasification processes: insight into catalyst development,
silicate with quantum chemistry calculations, Cem. Concr. Res. 24 (7) (1994) Process Safety and Environmental Protection 149 (169–184) (2021), https://doi.
1311–1316, https://doi.org/10.1016/0008-8846(94)90116-3 org/10.1016/j.psep.2020.10.047
[59] W.H. Casey, H.R. Westrich, Control of dissolution rates of orthosilicate minerals by [63] X. Bao, M. He, Z. Zhang, X. Liu, Crystal Structure and Some Thermodynamic
divalent metal–oxygen bonds, Nature 355 (6356) (1992) 157–159, https://doi.org/ Properties of Ca7MgSi4O16-Bredigite, Crystals 11 (2021) 14, https://doi.org/10.
10.1038/355157a0 3390/cryst11010014
[60] R. Caracas, X. Gonze, Ab initio determination of the ground-state properties of [64] S.V. Dimitrova, I.K. Mihailova, V.S. Nikolov, D.R. Mehandjiev, Adsorption capacity of
Ca2MgSi2O7 åkermanite, Phys. Rev. B 68 (18) (2003) 184102, https://doi.org/10. modified metallurgical slag, Bulgarian Chemical Communications 44 (2012) 30–36.
1103/PhysRevB.68.184102 [65] S. Saburi, A. Kawahara, C. Henmi, I. Kusachi, K. Kihara, The refinement of the
[61] C.J. Yu, J. Kundin, S. Cottenier, H. Emmerich, Ab initio modeling of glass corrosion: crystal structure of cuspidine Mineralogical Journal 8 (5) (1975) 286 298, https://
hydroxylation and chemisorption of oxalic acid at diopside and akermanite doi.org/10.2465/minerj.8.286

10

You might also like