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

Construction and Building Materials 276 (2021) 122235

Contents lists available at ScienceDirect

Construction and Building Materials


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

Accelerated carbonation of ladle furnace slag and characterization


of its mineral phase
Yuan-Rong Yi a,b,c,⇑, Yue Lin a, Yun-Cong Du a, Shu-qi Bai a, Zhong-le Ma a, Yin-guang Chen d
a
College of Resources and Environmental Sciences, Xinjiang University, Urumqi 830046, China
b
Key Laboratory of Oasis Ministry of Education, Xinjiang University, Urumqi 830046, China
c
Key Laboratory of Smart City and Environmental Modeling Autonomous Region, Urumqi 830046, China
d
College of Environmental and Energy Engineering, Tongji University, Shanghai 200092, China

h i g h l i g h t s

 The optimal parameters are 88 lm, 1:5, 20 °C, 500 ml/min, 700 rpm.
 The reaction was converted to laterites by reaction for 45 min.
 After 70 min of reaction, the reactants are converted into calcite crystal form.
 Quantitative analysis of mineral phase changes.

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

Article history: Ladle furnace slag is a high-calcium-based alkaline industrial solid waste inefficiently stacked at the site,
Received 19 July 2020 which consumes large amount of land and pollutes water resources. Carbonation of refining slag not only
Received in revised form 4 December 2020 solves the problem of poor compactness and poor stability of the refined slag as a product, but also elim-
Accepted 30 December 2020
inates waste and reduces greenhouse gases. Experiments were conducted to explore factors that affect
the carbonation of refining slag. The optimal reaction conditions were found to be temperature of
20 °C, liquid-solid ratio of 5, aeration rate of 600 ml/min, and stirring speed of 700 rpm. The slag with
Keywords:
less than 88 lm particle size showed the best carbonation effect with 39.6% degree of carbonation.
LFS
Carbonation
XRD, SEM, FT-IR, and DG-DTA methods were used to investigate the changes in mineralogical character-
Mineral phase change istics with time in the carbonation process of refining slag. Variations in mineral phase with time were
Quantitative analysis analyzed by semi-quantitative relative intensity ratio (RIR) method. TGA-DTG analysis quantified the
Mineral phase change in time series Ca conversion, CO2 absorption, and also estimated the CaCO3 mass at different time periods.
Ó 2021 Elsevier Ltd. All rights reserved.

1. Introduction liquid–solid ratio, etc.) [1,2]. The main reaction is the hydrolysis of
CO2 and the reaction of alkaline cations ionized by metal oxides to
With rapid developments of industries globally, issues relating form stable carbonates, among which CaO and MgO are the most
to carbon emissions have gained a lot of attention. Accelerated car- promising metal oxides in the carbonation reaction. The most
bonation process involves the sealing of CO2 in carbonate, which is desirable source of CaO and MgO is alkaline industrial solid waste
currently a safe and effective way of carbon sequestration. Due to such as: ladle furnace slag (LFS).
chemical stability and safe storage of carbonates, it is not only The mineral phase of refining slag is generally composed of var-
environmental friendly, but also reduces the greenhouse effect. ious oxides and silicates of Ca, Mg, and Al. It is generally from
Accelerated carbonation is a method for increasing the carbonation f-CaO, C2S(Ca2SiO4), C3S(Ca3SiO5), CS(CaSiO3), and silicate mineral
efficiency significantly by simulating natural weathering hydrate and other components [3–7]. These calcium mineral
conditions and artificially changing the process parameters phases may react with CO2. The reaction equation is as follows:
(rotation speed, gas flow rate, pressure, temperature, reaction time,
CaO(s) + CO2(g) () CaCO3(s) (DH =  179 kJ/mol) ð1Þ

⇑ Corresponding author at: College of Resources and Environmental Sciences, MgO(s) + CO2(g) () MgCO3(s) (DH =  118 kJ/mol) ð2Þ
Xinjiang University, Urumqi 830046, China.
E-mail address: yyrhyw@163.com (Y.-R. Yi).

https://doi.org/10.1016/j.conbuildmat.2020.122235
0950-0618/Ó 2021 Elsevier Ltd. All rights reserved.
Yuan-Rong Yi, Y. Lin, Yun-Cong Du et al. Construction and Building Materials 276 (2021) 122235

Ca2SiO4(s) + 2CO2(aq) + H2O(l) ! 2CaCO3(s) + SiO2H2O 2.2. Hydration carbonation experimental protocol
ð3Þ
The experiment is a solid-liquid-gas three-phase reaction, and
the equipment chooses to use a magnetic suspension heating stir-
Ca3SiO5(s) + 3CO2(aq) + H2O(l) ! 3CaCO3(s) + SiO2H2O
rer for heating and stirring, the temperature range of heating is
ð4Þ set/controlled at 20-80℃. CO2 with a flow rate of 99.9% controlled
by a gas flow meter is passed into the reaction space, measureing
Ca12Al14O33(s) + 12CO2(aq) + 7H2O(l) ! 12CaCO3(s) + 7Al2O3H2O the mass once every five minutes and recording the change. When
ð5Þ the mass no longer changes within 20 min, the reaction is deemed
to be complete. After the reaction was complete, the completed
Ca3Al2Si3O12(s) + 3CO2(aq) + 3H2O(l) ! 3CaCO3(s) slurry was subjected to solid-liquid separation, and the solid part
ð6Þ was dried to remove the influence of free water at 105 °C and char-
þ 3SiO2  H2O þ Al2O3
acterization analysis (XRD, SEM, FT-IR, TGA-DTG) was performed.

Ca(OH)2(s) + H2O(l) + 2CO2(aq) ! CaCO3(s) + 2H2O ð7Þ 2.3. Identification and analysis of minerals
LFS is the steel slag with the highest calcium content, used to
leach metal into solution [8,9], modified acid soil [10,11], for adsorp- The refining slag was ground and scanned using X-ray diffrac-
tion of heavy metals in water [12], etc. Its high calcium characteris- tometer (Bruker D8 advance, XRD, Germany) with Cu as an anode
tics make it more suitable for accelerated carbonation capture material at a scanning rate of 2°/min and 2h range of 10-80°. Scan-
materials. Bonenfant [13] analyzed the carbon sequestration capac- ning electron microscopy (LEO, 1430VP, SEM) was used to study
ity of LFS and electric arc furnace slag (EAFS) and found that the car- the micromorphology of the refining slag during the reaction.
bon sequestration capacity of LFS is 14 times more than that of EAFS. Infrared spectrometer (Bruker vertex 70, Germany) was used to
This indicated that LFS is a more ideal carbonation material in steel compare the structural changes before and after reaction, in the
slag. Researchers at home and abroad agree that free calcium oxide spectral range of 4000–400 cm1.
(f-CaO) in steel slag is a factor that causes poor stability of steel slag,
Wang [14] evaluated that the free calcium oxide content in the 2.4. Quantification of calcium carbonate
refined slag after carbonation is significantly reduced, and it has
higher stability and stability when used as a product. Therefore, Using semi-quantitative reference intensity ratios (RIR) method
hydration carbonation provides a broader research prospect for the [20], changes in phase composition of the refining slag were ana-
preparation of building materials from refined slag. lyzed by XRD. According to the ‘‘adiabatic” principle, when there
Although scholars have put forth the feasibility analysis of are N phases in the system, the mass fraction of the i-th phase
accelerated carbonation reaction of LFS, the carbonation reaction can be given by the additional RIR value of the JCPDS card and
process and the main of characteristics mineral phase change in the integral strength of each phase (Ii), as shown in formula (1):
the reaction are little known [15–17]. The carbonation mechanism
Ii =RIRi
has not fully been elucidated, but few articles have quantitatively W i ¼ PN ð1Þ
analyzed the quality of mineral phase at each stage [18,19]. There- i¼1 Ii =RIRi

fore, research on changes in carbonation mineral phase of LFS with The thermogravimetric analysis (TGA, Hitachi STA7300, Japan)
time series and quantitative analysis can more intuitively and was used to quantitatively analyze the changes in composition in
clearly show changes in mineral phase at each time period, and the carbonation reaction. The sample was heated from 10 °C to
provide more theoretical knowledge of mechanism of carbonation 1000 °C at average heating rate of 10 °C/min in N2 atmosphere.
reaction of steel slag or other alkaline industrial solid wastes. This helped to estimate the degree of carbonation, CO2 absorption,
In this paper, the influence of particle size, liquid-solid ratio, rota- CaCO3 quality, and the resulting mass loss.
tional speed, temperature, and gas flow rate on the carbonation reac-
tion was studied comprehensively, and the optimal parameters were
3. Results and discussion
determined. The changes in mineral phases during carbonation at
different time periods were investigated by X-ray diffractometry
3.1. Composition of refining slag
(XRD), infrared spectroscopy (FT-IR), thermogravimetric analysis
(DG-DTA), and scanning electron microscopy (SEM). Based on these
Table 1 presents the compositions of refining slags (LFS), basic
studies, the absorption of CO2, mass of CaCO3 and mass loss due to
oxygen slags (BOFS), and electric arc furnace slags (EAFS) from dif-
heat absorption during carbonation were quantitatively analyzed.
ferent countries. From the table, it is clear that the refining slag
This study provides further knowledge of the carbonation mecha-
constituted 40–60% of all the slags, with the highest calcium con-
nism of refined slag and understanding of the changes in mineral
tent amongst all steel slags. This is due to the large amount of lime-
phase composition in the carbonation process of refined slag.
stone used as a slag-forming agent in the refining process. BOFS
and EAFS use scrap steel and pig iron as raw materials, due to
2. Materials and methods which the Fe content and hardness are high. It can be seen from
the table that the calcium contents of these two slags are less.
2.1. Experimental materials Therefore, LFS is considered to show one of the best carbon capture
properties in steel slag. As shown in Fig. 1, three kinds of steel slag
The material used in the experiments was the waste residue and other solid wastes (MSWI Fly Ash) [21], Bayer red mud [22],
produced in the LF refining furnace (China, Xinjiang). The LF slag asbestos tailings [23], fly ash from burning of coal (Coal Fly Ash)
was pulverized (German Retsch jaw crusher) into granules and [24] were analyzed and compared, As shown in Fig. 1, high calcium
ground (Retsch RS200, Germany). The particles of different sizes and magnesium characteristics of LFS are found to be one of the
were selected and the experiments were carried out. The chemical most potential carbon capture materials in solid waste [25].
composition of refining slag was determined using an X-ray fluo- The theoretical carbon sequestration rate (Eq. (2)) proposed by
rescence spectrometer (PANalytical, XRF, Netherlands). Huntzinger [19] is used to estimate the theoretical amount of CO2
2
Yuan-Rong Yi, Y. Lin, Yun-Cong Du et al. Construction and Building Materials 276 (2021) 122235

Table 1
Comparison of slag used in this study with other slags from different countries.

Waste group Concentration


CaO Al2O3 SiO2 MgO SO3 FeOtotal F Na2O TiO2 K2O Cl P2O5 MnO Cr2O3
LFS This research 51.86 20.12 12.10 5.92 5.57 1.74 1.26 0.89 0.24 0.10 0.07 0.07 0.05 0.02
(China)
Canada [13] 58.1 4.6 26.4 6.2 0.03 4.3 – – 0.74 0.03- – 0.05 2.18 –
Spain [16] 50–58 4–19 12–20 7–12 0.9–1.5 1–4 – 0.03–0.07 0.1–1 0.01–0.02 – <0.01 0.3–0.6 0.01–0.1
Sweden [26] 42.5 22.9 14.2 12.6 2 0.0084 0.2 0.027
Croatia [27] 48.37 14.3 15 15.25 1.54 0.43 0.2 0.36 2.73 0.92
BOFS South Africa [28] 38–42 2–8 13–19 7–10 16–27 0.2–1.5 1–5
China [29] 42.42 1.61 11.04 7.19 0.11 27.37 0.5 2.09 2.28 0.15
Sweden [26] 45 1.9 11.1 9.6 23.9 0.083 0.02 3.1
Korea [30] 44.63 14.87 4.67 20.17 2.3 2.17
EAFS Canada [31] 35.23 10.78 9.41 9.77 24.22 0.03 less than0.01 3.84
Finland [3] 40.8 8.36 26.6 7.21 0.092 1.59 0.11 2.64 2.29 5.07
USA [32] 32.1 8.6 19.4 4.3 26.4 0.3 0.8
Spain [33] 24.4 12.21 15.35 2.91 – 34.36 – 0.19 0.56 1.52 – 1.19 5.57 0.99
Sweden [26] 38.8 6.7 14.1 3.9 26.5 0.024 0.02 5.0 0.268

Fig. 1. Three-phase diagram of refining slag and other industrial solid waste. Fig. 2. Effect of solid-liquid ratio on carbonation conversion rat.

captured by refining slag. It can be assumed that all of CaO is con- slag were continuously carbonated to form a stable mineral phase
verted to CaCO3, MgO is converted to MgCO3, and Na2O3 and K2O containing CaCO3 with a high degree of hardness. In addition, when
form Na2CO3 and K2CO3. In this study, the theoretical carbon the solid-liquid ratio was relatively high, the refining slag occupied
sequestration per kg of LFS reached 0.4474 kg CO2. a large amount of reaction space, which made it difficult for CO2 to
enter the reaction system. The gas permeability was reduced,
%ThCO2 ¼ 0:785ð%CaO  0:56%CaCO3  0:7%SO3 Þ% which limited the reaction process of hydration and carbonation
þ 1:091MgO þ 0:71Na2 O þ 0:468K2 O ð2Þ to some extent.
Fig. 3 shows the effect of particle size on carbonation reaction.
When the particle size was less than 88 lm, the carbonation
3.2. Optimal parameters for carbonation amount reached maximum value of 106 mg g1, and the equilib-
rium was attained at 130 min. When the particle size of reaction
Fig. 2 shows the effect of solid-liquid ratio on the carbonation material was 250 lm, the reaction rate and the carbonation effi-
reaction. It can be seen that increase in carbonation reached a max- ciency were the lowest. This was due to the fact that the reaction
imum value at solid-liquid ratio of 1:5. When the solid-liquid ratio occurred on the surface of the particles. The adhesion of a dense
was 1:9, the amount of refining slag was less, and resistance to the layer of Ca-O and Si-O prevented the leaching of the solute Ca2+.
leaching of Ca ions on the liquid film side of the gas–liquid inter- Scholars such as Du Yuncong [34] and Liang Xiaojie [35] proved
face was reduced. This resulted in an increase in the reaction rates that the leaching of Ca2+ continued to decrease as carbonation
of the gas and liquid in the entire unit reaction system and so the reaction progressed, indicating that carbonation prevented the
reaction reached saturation at 40 min. There was no leaching of leaching of Ca2+.
alkali metal in the solute and it combined with CO2 3 and HCO3
Aeration had an effect on the carbonation reaction, as seen from
to form a precipitate. According to the XRD analysis, the mineral Fig. 4. At 300–500 ml/min, carbonation extent increased with the
phase components such as C2S and Ca3Al2(SiO4)3 in the refining increase in aeration. At 500 ml/min, the carbonation was maximal.

3
Yuan-Rong Yi, Y. Lin, Yun-Cong Du et al. Construction and Building Materials 276 (2021) 122235

Fig. 3. Effect of particle size on carbonation conversion rate. Fig. 5. Effect of stirring speed on carbonation conversion rate.

With increase in aeration, the CO2 partial pressure continued to phase reactor and also increase in leaching rate of Ca2+. Increase
increase, which reduced the diffusion of CO2 and increased its par- in temperature caused instability of HCO 3 , which led to decrease

ticipation. This resulted in easier binding of CO2 to Ca2+ to form a in concentration of HCO 3 and consequently decrease in carbona-

stable carbonate precipitate. With further increase in aeration to tion. The optimum temperature for carbonation was found to be
600 ml/min, the lowest value of carbonation was observed. 20 °C. Selection of appropriate temperature helps in saving of
because the slower CO2 hydration reaction belongs to the speed energy (See Fig. 7).
control step and H2CO3 ionizes HCO 2
3 , CO3 faster, increasing the
Based on above experiments, refining slag with particle size less
ventilation, the CO2 cannot fully enter the reaction. The overflow than 88 lm, reaction temperature of 20 °C, solid-liquid ratio of 1:5,
in the system not only decreased the reaction efficiency, but also aeration rate of 500 ml/min and stirring speed of 700 rpm were
increased the participation of CO2 gas. found to be the best for carbonation of refining slag.
Stirring speed is one of the important parameters affecting dif-
fusion during the carbonation of refined slag. As seen from Fig. 5, at 3.3. Degree of carbonation
600–700 rpm carbonation was maximal. As the stirring speed was
increased further, the reaction slowed down and the equilibrium The method for quantitative analysis of carbonation efficiency is
time of reaction was reduced. Therefore, these results suggested based on the mass loss during carbonation at different tempera-
that the diffusion mass transfer resistance of Ca leaching and gas tures. Mass losses during carbonation reaction can be divided into
dissolution process in the refining slag carbonation is basically 4 stages: (1) Below 105 °C due to evaporation of free moisture (2)
eliminated at a rotation speed of 600 rpm. 105–350 °C due to dehydration of CSH (calcium silicate hydrate)
The relationship between temperature and carbonation extent and CAH (calcium aluminate hydrate) [36]. (3) Dehydration of Ca
is shown in Fig. 6. It can be seen that carbonation conversion and Mg hydroxide between 350 and 450 °C [37,38]. (4) Above
decreased with increase of reaction temperature. Increase in 450 °C due to heat loss from thermal decomposition of carbonate
temperature led to decrease in solubility of gas in the three- with release of CO2. Formula for calculation of total weight loss
is as follows:

Fig. 4. Effect of aeration on carbonation conversion rate. Fig. 6. Effect of temperature on carbonation conversion rate.

4
Yuan-Rong Yi, Y. Lin, Yun-Cong Du et al. Construction and Building Materials 276 (2021) 122235

Ca for each time period is shown in Table 3, and the carbonation


conversion was 39.6% in two hours. The peak at 650 °C increased,
indicating significant heat absorption, since large amount of heat
was required for thermal decomposition of CaCO3. Above 750 °C,
the differential thermal curve of refining slag had no further
changes, showing the various forms of refining slag after carbona-
tion. The carbonate was completely decomposed.

3.4. Characterization of carbonized products at different time periods


by XRD

For the refining slag that was carbonated for 5 min, the diffrac-
tion peak at 2h = 30–35°, corresponding to main phase of C2S
hydrate, in the XRD pattern (Fig. 8), changed with time. The ‘‘con-
vex bulge” finally disappeared, indicating that the LF refining slag
first contained water. When the reaction was carried out for
45 min, a characteristic peak of CaCO3 (vaterite) appeared. This
was consistent with the findings of Sauman [39], who suggested
Fig. 7. TG-DTG chart before and after reaction (Dp ~ 88 lm; L/S = 5 ml g1; T = 20℃; that carbonization of CSH produced incompletely crystalline and
500 ml min1; 700 rpm). unstable calcium carbonate crystal-vaterite. After 70 min of reac-
tion, no C2S phase was observed in the spectrum, indicating that
carbonization produced CaCO3. The main phase of the product after
Total weight lossðwt%Þ ¼ ðwt105 C  wt950 C Þ  100=wt105 C ð3Þ
carbonation for 2 h was CaCO3 (CaCO3, JCPDS: 085-0849), and the
In TGA measurements, the mass fraction of CO2 produced by intensity of diffraction peak was significant, since the large number
thermal decomposition of CaCO3 above 450 °C was used to repre- of Ca groups in the refining slag, such as f-CaO, CS, C2S, and Ca12-
sent the mass fraction of CO2 [2] as shown in Eq. (4): Al14O33 reacted with CO2. It formed an unstable crystalline vaterite,
CO2 ðwt%Þ ¼ ðDw1  Dw2Þ450950 C  100=wt105 C ð4Þ which eventually transformed into stable calcite and aragonite
crystal forms, with increase in carbonation time. This was consis-
ðCO2 ðwt%Þ=ð100  CO2 ðwt%ÞÞ  ð1=MWCO2 ðkg=molÞ tent with the results of Zhan [40], who attributed the peak at
fCa% ¼ ð5Þ 2h = 29.4° to the calcite form of CaCO3 crystal. Hence, the final main
CaOtotal ðwt%Þ=MWCaoðkg=molÞ
mineral phase for carbonation of the refining slag was calcite and
In Eq. (5), fCa% represented the Ca conversion rate, i.e. the was accompanied by needle shaped or plate-shaped aragonite
degree of carbonization; CaOtotal was the mass fraction of CaO crystal form [41]. Same conclusions were drawn from the semi-
before re-reaction; and MWCO2 (kg⁄(mol)) and MWCaO ((kg)/ quantitative analysis of carbonation products at different time
(mol)) were the molar masses of CO2 and CaO, respectively. The periods according to the RIR method. As shown in Table 2, the
TG-DTG plots and the calculations from Eqs. (3) and (4) showed calcium-based materials which reacted for 2 h were basically con-
that the total weight loss during carbonation process was 22.64%, verted into stable CaCO3 and Ca12Al14O33 and reaction was basi-
whereas the total CaCO3 loss was 13.35%. The conversion rate of cally complete after 15 min. The diffraction peak of Ca3Al2Si3O12

Fig. 8. XRD patterns at different times (Dp ~ 88 lm; L/S = 5 ml g1; T = 20 °C; 500 ml min1; 700 rpm).

5
Yuan-Rong Yi, Y. Lin, Yun-Cong Du et al. Construction and Building Materials 276 (2021) 122235

Table 2
Quantitative analysis of carbonation products at different times by RIR.

Major phrase JCPDS Fresh 5 15 45 70 End


Lime CaO 82-1691 8.54
Wollastonite CS 75-1396 27.66
Larnite C2S 70-0388 4.62 10.07 25.2 23.17 14.21
Quartz SiO2 82-1573 16.41 70.44 62.26 50.23 64.64 52.7
Vaterite CaCO3v 33-0268 0.86 1.41
Calcium CaCO3 85-0849 8.01 10.57 38.34
Grossular Ca3Al2Si3O12 72-1491 23.73 9.82 6.75 14.92 7.22 4.98
Mayenite Ca12Al14O33 09-0413 4.19 3.02 1.48
Magnetite Fe3O4 79-0416 14.84 5.79 2.9 3.68 3.36 3.97

Table 3
Thermal decomposition data, CO2 absorption and estimated CaCO3 quality of refining slag/%.

CO2 curing duration 45 min 70 min End Compared with other LFS
Degree of carbonation 7.4 12.9 39.6 10 [32]
Estimate the CO2 absorption of LF slag 3.1 5.6 17.7 1.9 [18]
Estimated quantity of CaCO3 formed in the carbonated pastes 6.5 11 29

(calcium aluminum garnet) reduced in intensity after 70 min reac- surrounded by inert SiO2 to form an external SiO2 enriched zone
tion. This indicated that Ca3Al2Si3O12 had very slow hydration and as shown in Fig. 9. MgO is also involved in the reaction. Since the
carbonation rates at normal temperature and pressure and was a characteristic peaks of CaCO3 and MgCO3 overlap, their individual
relatively stable mineral phase component. The diffraction peak peaks could not be distinguished from each other. The chemical
at 2h = 18° and 30° corresponded to Fe3O4, which showed no obvi- composition and carbon capture ability show that CaO was the
ous changes before and after the reaction, indicating that Fe did not main species that could effectively improve the carbonation reac-
participate in the reaction. tion ability.
The characteristic peaks of f-CaO appeared at 2h = 32.429°, Fig. 10 shows the morphological changes in refining slag at dif-
37.621°, and 54.258°. It can be seen that the diffraction peak ferent times of carbonation. The CaCO3 crystal included three crys-
decreased significantly after carbonization for 5 min, and there tal forms-calcite (diamond), aragonite (needle), and vaterite
was no diffraction peak for f-CaO after 70 min. Meanwhile, the (polycrystalline sphere) [44]. After 5 min of carbonization, poly-
characteristic peak for CaCO3 remained intact with increase in crystalline spherical CaCO3 could be identified, but the larger edge
reaction time. The product of carbonization of f-CaO was CaCO3. of the pores also showed a layer of CSH gel. With prolongation of
This suggested that f-CaO in the slag could be reduced in the car- reaction time, the carbonation products appeared mostly floccu-
bonization reaction. Hence, the problem of poor stability of the lated and were arranged more compactly. This could be due to
building materials prepared from the refining slag was resolved the fact that the gelation of Ca2SiO4 caused different crystal forms
[42]. The secondary phase of carbonation reaction was SiO2. of CaCO3 to bond together to form a dense coating, which increased
According to the study [43], the leaching of Ca resulted in a with- with reaction time. The pores were filled with the calcite crystal
drawing calcium silicate core surrounded by a Ca-depleted SiO2 form, which reduced the porosity and pore surface area. The SEM
zone and a formation of CaCO3-coating zone. This caused it to be image further showed dense calcium carbonate crystals formed

Fig. 9. Carbonation mechanism of LFS.

6
Yuan-Rong Yi, Y. Lin, Yun-Cong Du et al. Construction and Building Materials 276 (2021) 122235

Fig. 10. Topographical features after different time periods of reaction ((Dp ~ 88 lm; L/S = 5 ml g1; T = 20 °C; 500 ml min1; 700 rpm. (a) Fresh; (b) Carbonation for 5 min; (c)
Carbonation for 15 min (d) Carbonation for 70 min; (e) End.)

on the surface of refining slag. This hindered the leaching of Ca to prior to analysis, and there was no free water present in any of
the core and made it impossible to react with CO2 3 in the slurry, them. Hence, these peaks could be attributed to reabsorption of
eventually causing the reaction to stop. When the reaction was car- water after drying and CAOAH. Peak at 1424 cm1 after carbona-
ried out for 70 min, a layer of CaCO3 was superimposed that tion was ascribed to in-plane bending vibrations of CAOAH and
reduced the porosity. CO2
3 . The FT-IR spectra clearly showed significant changes in
Fig. 11 shows the FT-IR spectra of LFS carbonation products at absorption peak, before and after the reaction. Hence, it was evi-
different time periods. These results further confirmed the physic- dent that C-S-H and carbonate were produced. These findings were
ochemical characteristics of refining slag during carbonation and consistent with the XRD results. The absorption peaks at 465, 579,
were compared with XRD and DTG results. Peaks at 3642 cm1 and 851 cm1 seen in the spectrum before reaction, disappeared
and 3666 cm1 were due to OH stretching vibrations, before and after the reaction. A strong peak appeared at 876 cm1 for
after carbonation. All the samples in the experiment were dried carbonate after the reaction. At the same time, characteristic peaks

7
Yuan-Rong Yi, Y. Lin, Yun-Cong Du et al. Construction and Building Materials 276 (2021) 122235

mass difference due to carbonate to the mass at 950 °C. As in Eq.


(6), the mass loss of CaCO3 changed with the difference in mass
of carbonate before and after carbonation, as in Eq. (7):
Dw1
CO2 uptake ¼ ð6Þ
Weight at 950 C

100
Q uantity of CaCO3 ¼ ðDw1  Dw2Þ  ð7Þ
44
where Dw1 and Dw2 represented the weight loss after carbonation
and before carbonation at 550–850 °C, respectively. The following
table [6] shows the amount of heat loss in TGA, the amount of
CO2 absorbed and the mass of CaCO3 for each time period.

4. Conclusions

In this paper, it was found that the refining slag (less than
88 lm) with a temperature of 20 °C, a solid-liquid ratio of 1:5, a
Fig. 11. FT-IR diagram of time (Dp ~ 88 lm; L/S = 5 ml g1; T = 20 °C; 500 ml min1; aeration rate of 600 ml/min and rotation speed of 700 rpm had
700 rpm.) the best carbonation effect, and the carbonation degree reached
39.6%. f-CaO in the original sample of refining slag basically disap-
peared after 5 min of carbonation. When CS and C2S were carbon-
ated for 5 min, C-S-H gel was first generated, and spherical
of Fe-O and Si-O-Al observed at 524 cm1, indicated the presence aragonite was generated before 45 min of reaction. However, the
of aluminate and iron compounds, which was also consistent with more stable crystalline calcite and aragonite were generated after
the XRD results (See Fig. 12). 70 min of reaction. The reaction of Ca12Al14O33 was completed at
15 min, while the hydration carbonation of Ca3All2Si3O12 was slow,
3.5. TGA-DTG quantitative analysis and Ca3All2Si3O12 was a relatively stable mineral phase. The amor-
phous SiO2 did not show significant changes before and after the
Above 450 °C, the weight loss was generally due to the decom- reaction. As the carbonation time increased, an inert area was
position of carbonate, CO2 being the main gas released [40]. formed, which prevented the reaction to proceed. The absorption
According to XRD analysis, formation of vaterite started when of CO2 was 17.7% and 29% of CaCO3 was generated, according to
the carbonation reaction time was 5 min. This crystal form was dif- TGA-DTG.
ferent from the calcite, which had a better crystalline state. TGA Therefore, it could be concluded that refining slag is a potential
results further confirmed that the endothermic peak gradually carbon capture material. This method of carbonation could acceler-
increased with carbonation time. This demonstrated the transition ate the carbonation efficiency of refining slag. Particle size was one
of the reaction product to the calcite form, with better crystallinity. of the important factors that controlled the reaction rate. However,
From the TG and DTG plots, the continuous loss of mass confirmed smaller the particle size, the higher is the cost of its grinding. So
that C-S-H and CAH were formed during carbonation. Especially, our future research will evaluate methods to increase the carbon-
the DTG plot for 45-minute carbonation reaction showed signifi- ation efficiency with low energy consumption. In addition, the
cant mass loss between 105 and 350 °C, which was identified from change of mineral phase indicated that as the reaction time
XRD. The main mineral phase was an anastomosis of C2S hydrate. increased, the carbonation products gradually transformed into
The amount of CO2 absorbed could be expressed as the ratio of more stable crystals. In future, NMR and other techniques would

Fig. 12. TG-DTG curve of LFS carbonation.

8
Yuan-Rong Yi, Y. Lin, Yun-Cong Du et al. Construction and Building Materials 276 (2021) 122235

be used to further explore the changes in structures of mineral [20] Y. Zhang, W. Liu, X. Liu, Segregation behavior and evolution mechanism of
iron-rich phases in molten magnesium alloys, J. Mater. Sci. Technol. 32 (1)
phase of carbonation, so as to provide more information regarding
(2016) 48–53, https://doi.org/10.1016/j.jmst.2015.10.013.
the carbonation mechanism of refining slag. [21] E. Rendek, G. Ducom, P. Germain, Carbon dioxide sequestration in municipal
solid waste incinerator (MSWI) bottom ash, J. Hazard. Mater. 128 (1) (2006)
Declaration of Competing Interest 73–79, https://doi.org/10.1016/j.jhazmat.2005.07.033.
[22] Y. Yi, M. Han, Y.u. Li-an, Reaction characteristics of CO2 captured by red mud,
CIESC. J. 62 (9) (2011) 2635–2642, https://doi.org/10.3969/j.issn.0438-
The authors declare that they have no known competing finan- 1157.2011.09.037.
cial interests or personal relationships that could have appeared [23] F. Larachi, I. Daldoul, G. Beaudoin, Fixation of CO2 by chrysotile in low-pressure
dry and moist carbonation: Ex-situ and in-situ characterizations, Geochim.
to influence the work reported in this paper. Cosmochim. Acta 74 (11) (2010) 3051–3075, https://doi.org/10.1016/j.
gca.2010.03.007.
Acknowledgement [24] A. Uliasz-Bocheńczyk, E. Mokrzycki, Z. Piotrowski, R. Pomykała, Estimation of
CO2 sequestration potential via mineral carbonation in fly ash from lignite
combustion in Poland, Energy Procedia 1 (1) (2009) 4873–4879, https://doi.
The study was funded by the China Natural Science Foundation org/10.1016/j.egypro.2009.02.316.
(51868072). [25] D. Bonenfant, L. Kharoune, S. Sauvé, R. Hausler, P. Niquette, M. Mimeault, M.
Kharoune, Molecular analysis of carbon dioxide adsorption processes on steel
slag oxides, Int. J. Greenhouse Gas Control 3 (1) (2009) 20–28, https://doi.org/
References‘ 10.1016/j.ijggc.2008.06.001.
[26] M. Tossavainen, F. Engstrom, Q. Yang, N. Menad, M. Lidstrom Larsson, B.
[1] W.J.J. Huijgen, R.N.J. Comans, Mineral CO 2 sequestration by steel slag Bjorkman, Characteristics of steel slag under different cooling conditions,
carbonation, Environ. Sci. Technol. 39 (24) (2005) 9676–9682, https://doi. Waste Manage. 27 (10) (2007) 1335–1344, https://doi.org/10.1016/j.
org/10.1021/es050795f.s001. wasman.2006.08.002.
[2] E.-E. Chang, S.-Y. Pan, Y.-H. Chen, C.-S. Tan, P.-C. Chiang, Accelerated [27] A. Radenović, J. Malina, T. Sofilić, Characterization of ladle furnace slag from
carbonation of steelmaking slags in a high-gravity rotating packed bed, J. carbon steel production as a potential adsorbent, Adv. Mater. Sci. Eng. 2013
Hazard. Mater. 227-228 (2012) 97–106, https://doi.org/10.1016/j. (2013) 1–6, https://doi.org/10.1155/2013/198240.
jhazmat.2012.05.021. [28] F.J. Doucet, Effective CO2-specific sequestration capacity of steel slags and
[3] S. Teir, S. Eloneva, C.-J. Fogelholm, R. Zevenhoven, Dissolution of steelmaking variability in their leaching behaviour in view of industrial mineral
slags in acetic acid for precipitated calcium carbonate production, Energy 32 carbonation, Miner. Eng. 23 (3) (2010) 262–269, https://doi.org/10.1016/j.
(4) (2007) 528–539, https://doi.org/10.1016/j.energy.2006.06.023. mineng.2009.09.006.
[4] R. Baciocchi, G. Costa, M. Di Gianfilippo, A. Polettini, R. Pomi, A. Stramazzo, [29] L. Mo, F. Zhang, M. Deng, F. Jin, A. Al-Tabbaa, A. Wang, Accelerated carbonation
Thin-film versus slurry-phase carbonation of steel slag: CO2 uptake and effects and performance of concrete made with steel slag as binding materials and
on mineralogy, J. Hazard. Mater. 283 (2015) 302–313, https://doi.org/10.1016/ aggregates, Cem. Concr. Compos. 83 (2017) 138–145, https://doi.org/10.1016/
j.jhazmat.2015.09.016. j.cemconcomp.2017.07.018.
[5] C. Shi, S. Hu, Cementitious properties of ladle slag fines under autoclave curing [30] M.-W. Choi, S.-M. Jung, Crystallization behavior of melted BOF slag during
conditions, Cem. Concr. Res. 33 (11) (2003) 1851–1856, https://doi.org/ non-isothermal constant cooling process, J. Non-Cryst. Solids 468 (2017) 105–
10.1016/S0008-8846(03)00211-4. 112, https://doi.org/10.1016/j.jnoncrysol.2017.04.042.
[6] L. Mo, F. Zhang, M. Deng, Mechanical performance and microstructure of the [31] M. Mahoutian, Y. Shao, A. Mucci, B. Fournier, Carbonation and hydration
calcium carbonate binders produced by carbonating steel slag paste under CO2 behavior of EAF and BOF steel slag binders, Mater. Struct. 48 (9) (2015) 3075–
curing, Cem. Concr. Res. 88 (2016) 217–226, https://doi.org/10.1016/j. 3085, https://doi.org/10.1617/s11527-014-0380-x.
cemconres.2016.05.013. [32] S.N. Lekakh, C.H. Rawlins, D.G.C. Robertson, V.L. Richards, K.D. Peaslee, Kinetics
[7] C. Shi, Characteristics and cementitious properties of ladle slag fines from steel of aqueous leaching and carbonization of steelmaking slag, Metall. Mater.
production, Cem. Concr. Res. 32 (3) (2002) 459–462, https://doi.org/10.1016/ Trans. B 39 (1) (2008) 125–134, https://doi.org/10.1007/s11663-007-9112-8.
S0008-8846(01)00707-4. [33] M.P. Luxán, R. Sotolongo, F. Dorrego, E. Herrero, Characteristics of the slags
[8] Y. Zhou, L. Wu, J. Wang, H. Wang, Y. Dong, Alumina extraction from high- produced in the fusion of scrap steel by electric arc furnace, Cem. Concr. Res.
alumina ladle furnace refining slag, Hydrometallurgy 140 (2013) 14–19, 30 (4) (2000) 517–519, https://doi.org/10.1016/S0008-8846(99)00253-7.
https://doi.org/10.1016/j.hydromet.2013.08.007. [34] D.u. Yuncong, Y. Yi, Z. Ma, S. Bai, M. Fang, W. Ma, Change of Ca occurrences
[9] S.-W. Lee, Y.-J. Kim, J.-H. Bang, S. Chae, CaCO3 film synthesis from ladle furnace state in carbonation process of LF refining slag, J Iron. Steel. Res. Int. 32 (03)
slag: morphological change, new material properties, and Ca extraction (2020) 195–203, https://doi.org/10.13228/j.boyuan.issn1001-0963.20190161.
efficiency, Int. J. Miner Metall Mater. 25 (12) (2018) 1447–1456, https://doi. [35] Xiaojie Liang, Studies on Relationship of Carbonation with Hydration of Steel
org/10.1007/s12613-018-1699-z. Slag and Cement.
[10] I. Anger, E. Volceanov, G. Plopeanu, L. Ilie, L.G. Popescu, G.A. Moise, Impact of [36] V. Ortega-López, J.M. Manso, I.I. Cuesta, J.J. González, The long-term
metallurgical ladle furnace slag on acid soil remediation in agriculture, Rev. accelerated expansion of various ladle-furnace basic slags and their soil-
Rom. Mater. 48 (1) (2018) 76–82. stabilization applications, Constr. Build. Mater. 68 (2014) 455–464, https://doi.
[11] J.M. Manso, V. Ortega-López, J.A. Polanco, J. Setién, The use of ladle furnace slag org/10.1016/j.conbuildmat.2014.07.023.
in soil stabilization, Constr. Build. Mater. 40 (2013) 126–134, https://doi.org/ [37] M.N. Moliné, W.A. Calvo, A.G.T. Martinez, P.G. Galliano, Ambient weathering of
10.1016/j.conbuildmat.2012.09.079. steelmaking ladle slags, Ceram. Int. 44 (15) (2018) 18920–18927, https://doi.
[12] A. Radenović, G. Medunić, T. Sofilić, The use of ladle furnace slag for the org/10.1016/j.ceramint.2018.07.128.
removal of hexavalent chromium from an aqueous solution, Metall. Res. [38] J. Waligora, D. Bulteel, P. Degrugilliers, D. Damidot, J.L. Potdevin, M. Measson,
Technol. 113 (6) (2016) 606, https://doi.org/10.1051/metal/2016040. Chemical and mineralogical characterizations of LD converter steel slags: a
[13] D. Bonenfant, L. Kharoune, Sébastien Sauvé, R. Hausler, P. Niquette, M. multi-analytical techniques approach, Mater. Charact. 61 (1) (2010) 39–48,
Mimeault, M. Kharoune, CO 2 sequestration potential of steel slags at ambient https://doi.org/10.1016/j.matchar.2009.10.004.
pressure and temperature, Ind. Eng. Chem. Res. 47 (20) (2008) 7610–7616, [39] Z. Šauman, Carbonization of porous concrete and its main binding
https://doi.org/10.1021/ie701721j. components, Cem. Concr. Res. 1 (6) (1971) 645–662, https://doi.org/10.1016/
[14] Y. Wang, P. Suraneni, Experimental methods to determine the feasibility of 0008-8846(71)90019-6.
steel slags as supplementary cementitious materials, Constr. Build. Mater. 204 [40] B.J. Zhan, D.X. Xuan, C.S. Poon, C.J. Shi, Mechanism for rapid hardening of
(2019) 458–467, https://doi.org/10.1016/j.conbuildmat.2019.01.196. cement pastes under coupled CO2-water curing regime, Cem. Concr. Compos.
[15] M. Mahoutian, Z. Ghouleh, Y. Shao, Carbon dioxide activated ladle slag binder, 97 (2019) 78–88, https://doi.org/10.1016/j.cemconcomp.2018.12.021.
Constr. Build. Mater. 66 (2014) 214–221, https://doi.org/10.1016/ [41] S. Kodama, T. Nishimoto, N. Yamamoto, K. Yogo, K. Yamada, Development of a
j.conbuildmat.2014.05.063. new pH-swing CO2 mineralization process with a recyclable reaction solution,
[16] J. Setién, D. Hernández, J.J. González, Characterization of ladle furnace basic Energy 33 (5) (2008) 776–784, https://doi.org/10.1016/j.energy.2008.01.005.
slag for use as a construction material, Constr. Build. Mater. 23 (5) (2009) [42] X. Yin, C. Zhang, J.i. Yang, B. Li, G. Wang, Research on stabilization of free CaO in
1788–1794, https://doi.org/10.1016/j.conbuildmat.2008.10.003. basic oxygen furnace slag with SiO2 bearing acidifier at high temperature,
[17] M. Uibu, R. Kuusik, L. Andreas, K. Kirsimäe, The CO2 -binding by Ca-Mg- Mater. Rev. 32 (2) (2018) 301–306, https://doi.org/10.11896/j.issn.1005-
silicates in direct aqueous carbonation of oil shale ash and steel slag, Energy 023X.2018.02.028.
Procedia 4 (2011) 925–932, https://doi.org/10.1016/j.egypro.2011.01.138. [43] C. Hall, D.J. Large, B. Adderley, H.M. West, Calcium leaching from waste
[18] D.N. Huntzinger, J.S. Gierke, S.K. Kawatra, T.C. Eisele, L.L. Sutter, Carbon dioxide steelmaking slag: Significance of leachate chemistry and effects on slag grain
sequestration in cement kiln dust through mineral carbonation, Environ. Sci. mineralogy, Miner. Eng. 65 (2014) 156–162, https://doi.org/10.1016/j.
Technol. 43 (6) (2009) 1986–1992, https://doi.org/10.1021/es802910z. mineng.2014.06.002.
[19] S.-Y. Pan, R. Adhikari, Y.-H. Chen, P. Li, P.-C. Chiang, Integrated and innovative [44] H. Cölfen, Precipitation of carbonates: recent progress in controlled production
steel slag utilization for iron reclamation, green material production and CO2 of complex shapes, Curr. Opin. Colloid Interface Sci. 8 (1) (2003) 23–31,
fixation via accelerated carbonation, J. Cleaner Prod. 137 (2016) 617–631, https://doi.org/10.1016/S1359-0294(03)00012-8.
https://doi.org/10.1016/j.jclepro.2016.07.112.

You might also like