You are on page 1of 7

This is an open access article published under an ACS AuthorChoice License, which permits

copying and redistribution of the article or any adaptations for non-commercial purposes.

Article

Cite This: ACS Omega 2019, 4, 3198−3204 http://pubs.acs.org/journal/acsodf

Interaction of Pristine Hydrocalumite-Like Layered Double


Hydroxides with Carbon Dioxide
Anand N. Narayanappa and P. Vishnu Kamath*
Department of Chemistry, Central College, Bangalore University, Bangalore 560 001, India
*
S Supporting Information

ABSTRACT: The layered double hydroxides (LDHs) of Ca2+ and


trivalent cations, Al3+ and Fe3+, are single-source precursors to
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

generate supported CaO, which picks up CO2 from the gas phase in
the temperature range 350−550 °C. The supports are ternary oxides,
mayenite, and Ca2Fe2O5. The uptake capacity of the Fe3+-containing
Downloaded via BANGALORE UNIV on February 14, 2019 at 07:18:07 (UTC).

LDH at 1.9 mmol g−1 is two times the capacity of the Al3+-
containing LDH. The product of CO2 uptake is calcite CaCO3. It is
observed that the intercalated chloride ions reduce the thermal
penalty by inducing the early decomposition of CaCO3. In the case
of the chloride-intercalated LDHs of Ca2+ and Fe3+, the CaCO3
formed is completely decomposed at 900 °C. This is in contrast with
the CaCO3 formed from bare CaO, which shows no sign of
decomposition at 900 °C under similar conditions. This work shows
that the hydrocalumite-like LDHs are candidate materials for CO2 mineralization.

1. INTRODUCTION for CO2 capture when compared to the sintered LDHs for the
following reasons.
The accumulation of anthropogenic CO2 in the atmosphere is
responsible for the rise in the temperature of the earth’s (i) The LDHs comprise positively charged metal hydroxide
atmosphere. There are major efforts underway to mineralize layers having the general formula [M-
CO2 and store it as a benign solid carbonate, an approach (II)1−xM′(III)x(OH)2]x+ (x = 0.2−0.33), wherein,
known as carbon capture and storage.1,2 An alternative is to use M(II) is commonly Mg2+ and M′(III) is commonly
the mineralized form as a feedstock for pure CO2 for other Al3+, Cr3+, and Fe3+. Anions are intercalated in the
interlayer region. The LDHs are structurally and
chemical processes.3,4 In the latter approach, the temperature
functionally the “inverse” of the cationic clays. While
difference between CO2 uptake and its eventual regeneration the cationic clays exhibit surface acidity, the LDHs are
constitutes the thermal penalty.5,6 For purposes of energy endowed with surface basicity.14 The interaction of
efficiency, materials with low thermal penalty are preferred. pristine LDHs with CO2 has the character of an acid−
A number of materials have been used as candidates for CO2 base reaction whereby the LDHs are expected to have a
capture and are reviewed elsewhere.7 Among them, sintered high affinity for CO2.
layered double hydroxides (LDHs) comprise a prominent (ii) Structurally, the LDHs have interlayer galleries which
family of materials. LDHs decompose to yield mixed metal provide space for storage. The LDHs are also known to
oxides (MMOs).8 Metal oxides in general exhibit surface swell and function as receptacles for increasing numbers
acidity and are not expected to mineralize CO2, as the latter is of intercalated species by means of a hydrogen-bonded
also acidic in nature. Nevertheless, a modest CO2 uptake network.15
capacity of 0.3 mmol g−1 has been reported for the MMOs of Despite these advantages, the pristine hydrotalcite-like
Mg2+ and Al3+ obtained by the thermal decomposition of [Mg−Al] LDH did not gain any mass when exposed to CO2
hydrotalcite-like LDHs.9 This capacity is attributed to the (pressure 1 atm) over the temperature range 30−400 °C,16
adsorption of CO2 within the morphological pores of the although the [Mg−Al] LDHs sorb carbonate anions from
MMO. The same MMO is also reported to exhibit a CO2 solution in quantities proportional to their layer charge.
uptake capacity of 0.53 mmol g−1.10 Modifications in the Another important candidate LDH for CO2 mineralization is
morphology of the crystallites and amorphization led to only the hydrocalumite-like LDH, in which M(II) is the Ca2+ ion
modest gains in the CO2 uptake capacity.11,12 LDHs infused and M′(III) is Al3+ or Fe3+.17 Although the conversion of
with up to 30% by weight of K2CO3 were found to show CO2
uptake capacity of ∼1 mmol g−1 of LDH after calcination.13 Received: January 10, 2019
From chemical, structural, and compositional criteria, Accepted: February 1, 2019
pristine LDHs offer themselves as superior candidate materials Published: February 13, 2019

© 2019 American Chemical Society 3198 DOI: 10.1021/acsomega.9b00083


ACS Omega 2019, 4, 3198−3204
ACS Omega Article

Table 1. Composition of the Hydrocalumite-Like LDHs


mass loss from TGA [%]
loss of H2O total loss
2+ 3+
Ca content M content anion content anion content
sample [wt. %]a [wt. %]a [wt. %]a [mol %]b N2 CO2 N2 CO2 approximate composition
[Ca−Al−Cl] 26.2 (28.8) 9.9 (9.7) 7.5 (7.63) 0.6 (1.0) 14.7 13.4 41.1 37.7 [Ca2Al(OH)6]Cl0.60(OH)0.40·2.3H2O
[Ca−Fe−Cl] 27.1 (26.1) 19.8 (18.8) 5.1 (5.1) 0.43 (1.0) 14.4 14.1 35.8 35.8 [Ca2Fe(OH)6]Cl0.43(OH)0.57·2.45H2O
[Ca−Al−NO3] 27.6 (27.4) 9.7 (9.2) 13.8 (12.7) 0.6 (1.0) 13.3 11.1 47.3 30 [Ca2Al(OH)6](NO3)0.60(OH)0.40·2.2H2O
a
Values in parenthesis are calculated based on the approximate composition given in the last column. bValues given in parenthesis are expected to
be the nominal composition of the LDHs (x = 0.33).

Ca(OH)2/CaO to CaCO3 is widely studied in the context of preference to the LDH during coprecipitation.21 The [Ca−M′]
deep well disposal of CO2 under high pressure and its LDHs are generally obtained with intercalated Cl− ions.22 The
subsequent mineralization to CaCO3,18 there is relatively less most common of these having the formula Ca2Al(OH)6Cl·
work on the sorption and uptake of CO2 by the [Ca−M′] (M′ 2H2O is called Friedel’s salt.17 The as-prepared [Ca−M′−Cl]
= Al3+ and Fe3+) LDHs. The [Ca−M′] LDHs are of interest (M′ = Al and Fe) LDHs are highly crystalline, and the powder
for many reasons in both the pristine and sintered forms. X-ray diffraction (PXRD) patterns (Figure S1 in the
(1) Given its high natural abundance, Ca2+ is ubiquitous in Supporting Information) exhibit sharp 00l (l = 3n) reflections
soil and offers economic advantages over Mg2+. in the low-angle region (5−25° 2θ) typical of layered materials
(2) Ca(OH)2 is more basic than Mg(OH)2 and is expected followed by hkl reflections which obey the reflection
to take up CO2 quantitatively in a facile manner to yield conditions −h + k + l = 3n (Table S1) in the high-angle
CaCO3 (expected uptake 13.5 mmol g−1). regions (2θ > 25°), indicative of rhombohedral symmetry. The
refined cell parameters are in agreement with earlier reports on
(3) The LDH of Ca and Al crystallizes in a cation-ordered
these LDHs.17,22 The [Ca−Al−NO3] LDH crystallizes in the
structure (space group R3̅), in which the Ca2+ ion has a
hexagonal symmetry (Figure S2, Table S2). Given the large
seventh coordination with a water molecule in the
difference in the ionic radii of Ca2+ (114 pm) and the trivalent
interlayer.17 In the hydrocalumite structure, the Ca2+ ion
cations (Al3+ 67.5 pm and Fe3+ 78.5 pm), the [Ca−M′] LDHs
is displaced away from the central plane of Al3+ ions.
crystallize in cation-ordered structures. The a-parameter (a ≈
This displacement facilitates coordination with a water
5.74 Å) is a measure of the M′···M′ distance within the metal
molecule in the interlayer. The puckered nature of the
hydroxide layer and is typical of the composition x = 0.33
metal hydroxide layer helps to trap the intercalated ion.
([Ca2+]/[M′] = 2).
For instance, when the [Ca−Al−Cl] LDH is dehydrated,
The approximate formulae of the LDHs (Table 1) were
the Cl− ion moves to the seventh coordination of Ca2+
arrived at by a combination of different independent
and is efficiently grafted to the layer. A similar
estimations. As an illustration, the methodology is described
phenomenon in the presence of CO32− could help
in detail for [Ca−Al−Cl]. The metal contents determined
immobilize CO2 more effectively than otherwise.
independently by a combination of AAS and gravimetry
Thereby, the [Ca−Al] LDH offers a different structure
correspond to the nominal composition [Ca2M′(OH)2]+. The
model for the mineralization of CO2 and its confinement
Cl− content estimated independently by ion chromatography
in the interlayer gallery.
was found to be substoichiometric. The shortfall of negative
(4) The decomposition temperature of Ca-containing LDHs charge required to neutralize the positive charge on the metal
is in the range of 300−400 °C. The flue gas temperature hydroxide layer was made up by the inclusion of hydroxyl ions
is very close to the decomposition temperature of Ca- in the interlayer to yield the LDH composition [Ca2Al-
containing LDHs facilitating the reaction between the (OH)6]Cl0.6(OH)0.4. Such an assumption is reasonable as the
LDH and CO2. precipitation is carried out at high pH (11.5−12). There is
(5) LDHs decompose with a mass loss of 30−40%. The evidence to suggest that LDHs comprising univalent anions are
large mass loss results in the formation of a highly rarely single anion materials and generally contain more than
porous MMO residue which is not only compositionally one anion.23,24 The intercalated water content was estimated
metastable19 but also has a large reactive surface area.20 by thermogravimetric analysis (TGA).
These conditions are ideally suited for the reconstruc- The TGA of the as-prepared [Ca−Al−Cl] LDH (Figure 1a)
tion of the LDH with the uptake of CO2 and water vapor carried out in flowing N2 matches with similar data published
from the flue gas. This makes Ca-containing LDHs by other authors.17 The mass loss occurs in three well-defined
interesting in CO2 sorption studies. sigmoidal steps, the inflection temperatures being 146 °C (step
As a first step to examining the suitability of the [Ca− I), 343 °C (step II), and 682 °C (step III) (Table 2).
M′(III)] LDHs as sorbents for CO2, this work reports the Assuming step I to correspond to dehydration yields a water
uptake of CO2 by a cohort of [Ca−M′−X] LDHs (M′ = Al3+ content corresponding to 2.3 water molecules per formula unit.
and Fe3+; X = Cl− and NO3−). The total mass loss (observed 41.1%; expected 41.4%)
corresponds to the complete decomposition reaction
2. RESULTS
[Ca 2Al(OH)6 ]Cl 0.6(OH)0.4 ·2.3H 2O → 2CaO + 0.5Al 2O3
A majority of the naturally occurring as well as laboratory-
synthesized LDHs have CO32− ions intercalated in the The mass loss associated with step II corresponds to
interlayer region. CO32−-intercalated [Ca−M′] LDHs are less dehydration and dehydroxylation (observed 16.2%; expected
known owing to the fact that the single cation CaCO3 forms in 20.6%). The third mass loss (step III) corresponds to anion
3199 DOI: 10.1021/acsomega.9b00083
ACS Omega 2019, 4, 3198−3204
ACS Omega Article

Table 3. Mass Change in TGA under Flowing CO2


step Ia step IIa step IIIb step IVa% step Vc
LDH % (°C) % (°C) % (°C) (°C) % (°C)
[Ca−Al−Cl] 13.4 14.8 4.6 14.5 8.7
(132) (380) (497) (800−869) (810)
[Ca−Al−NO3] 11.1 12.5 4.6 11.1
(123) (351) (467.5) (511−900)
[Ca−Fe−Cl] 13.9 12.1 8.5 18.3 (774)
(147) (322) (497)
a
Mass loss. bMass gain in heating cycle. cMass gain in cooling cycle.

are comparable to those obtained in the control under


flowing N2. However, following these two steps, unlike
in the control experiment, a plateau is established from
394 to 460 °C (plateau-I).
(ii) At the end of plateau-I, the LDH gains ∼4.5% mass (∼1
mmol g−1) in the temperature range 460−555 °C.
Thereafter, the mass is nearly constant with a downward
bias up to 750 °C (plateau-II). It is important to note
that the mass gain in flowing CO2 begins after the
dehydroxylation step. A prolonged stay at the plateau
temperature (T = 550 °C) in a separate experiment did
not lead to further mass gain (Figure S3).
(iii) A mass loss of ∼4.6% is observed at 788 °C.
(iv) On staying at 900 °C for up to 30 min, 37.7% total mass
loss is observed. This is less than the loss expected
(41.8%) for the complete decomposition of the LDH.
On cooling, the residue instantaneously regains some of
the mass earlier lost at T > 800 °C and reaches a
constant value at 700 °C (plateau-III). However,
plateau-III is nearly 5.5% mass lower than plateau-II
Figure 1. TGA and DTG curves of [Ca−Al−Cl] LDH in (a) flowing and ∼1% mass lower than plateau-I showing consid-
N2 and (b) flowing CO2.
erable hysteresis.
Table 2. Mass Loss Steps in TGA under Flowing N2 In separate experiments, the LDH was decomposed at the
temperatures corresponding to the three plateaus and
step II % step III % step IV % characterized by PXRD. The phase obtained at plateau-I
LDH step I % (°C) (°C) (°C) (°C)
(460 °C) has a reflection at ∼29.4° 2θ characteristic of 104
[Ca−Al−Cl] 14.7 (146) 16.2 (343) 10.2 (682) reflection of calcite CaCO3. The poor signal-to-noise ratio and
[Ca−Al−NO3] 13.3 13.2 (307) 12.5 (570) 8.3 (715) background are indicative of a significant volume fraction of X-
(118, 161)
[Ca−Fe−Cl] 14.4 (161) 12.1 (278) 9.4 (672)
ray amorphous content. The absence of the intense basal
reflections shows that the layered structure has collapsed. On
cooling back to the ambient, this phase does not regain mass,
expulsion (observed 10.2%; expected 6%). The difference in showing that dehydroxylation is not reversible. However, on
expected and observed values at the end of the steps II and III further heating in flowing CO2 to the temperature of plateau-II
arises because of overlapping of steps but the expected and (640 °C), CaCO3 formation is evident from the PXRD
observed total mass loss are in very good agreement. The step- pattern. The IR spectrum of the residue at plateau-II shows an
wise decomposition is enhancement of the absorption at 1409 cm−1 because of calcite
formation (see Figure S4 and the accompanying discussion of
[Ca 2Al(OH)6 ]Cl 0.6(OH)0.4 the IR spectra). At plateau-III, a mixture of CaCO3 and the
→ 2CaO + 0.6AlOCl + 0.2Al 2O3 ternary oxide mineral mayenite, Ca12Al14O33, is observed
(Figure 2). The thermodynamically stable ternary oxide is
→ 2CaO + 0.5Al 2O3 obtained at 900 °C and does not react with CO2. Only a part
A similar approach was used to arrive at the approximate of the CaO formed in excess over the ternary oxide at 900 °C
formulae of the other LDHs (Table 1). It is noteworthy that in gains CO2 on cooling to the temperature of plateau-III
the cooling cycle, the mass of the residue is unchanged, accounting for the hysteresis in the TGA.
showing that the decomposition is irreversible. Similar observations are made in the case of the [Ca−Fe−
The TGA measurement was repeated in flowing CO2 to Cl] LDH (Figure 3). The total mass loss (observed 35.8%;
examine the interaction of the LDH with CO2 in the gas phase expected 36.5%) in flowing N2 corresponds to the decom-
(Figure 1b, Table 3). The following observations were made: position reaction
(i) During the heating cycle, the LDH loses 28% of its mass [Ca 2Fe(OH)6 ]Cl 0.43(OH)0.57 ·2.45H 2O
up to 394 °C in two sigmoidal steps with inflection at
132 °C (step I) and 380 °C (step II). These two steps → 2CaO + 0.5Fe2O3

3200 DOI: 10.1021/acsomega.9b00083


ACS Omega 2019, 4, 3198−3204
ACS Omega Article

Figure 2. PXRD patterns of the residues obtained from the Figure 4. PXRD patterns of the residues obtained from the
decomposition of [Ca−Al−Cl] LDH at plateaus (a) P-I, (b) P-II, decomposition of [Ca−Fe−Cl] LDH at (a) P-I, (b) P-II, and (c)
and (c) P-III. Reflections marked with asterisks are due to calcite P-III. Reflections marked with asterisk correspond to CaCO3.
CaCO3. Unmarked reflections are due to mayenite. Unmarked reflections correspond to Ca2Fe2O5.

CaCO3 formed at plateau-II. The total mass loss corresponds


to a mixed oxide residue. On cooling, this oxide residue regains
mass in two minor steps (850 and 650 °C) to yield plateau-III
at which the residue corresponds to a mixture of CaCO3 and
Ca2Fe2O5. Evidently, the ternary oxide phase formed at 900 °C
does not pick up CO2 on cooling, resulting in a much greater
hysteresis between plateaus-I and III than that in the case of
the [Ca−Al] LDH.
As the Cl− ion is nonvolatile at moderate temperatures (T <
500 °C), similar experiments were performed on the NO3−-
intercalated LDH to see if the nitrate-LDH shows a higher
CO2 uptake than the chloride-LDH. Under flowing N2, the
[Ca−Al−NO3] LDH undergoes complete decomposition
(Figure S5, observed mass loss 47.3%; expected 44.8%).
Under flowing CO2, an uptake of 4.6% (∼1 mmol g−1 original
LDH) is observed (Figure 5) at plateau-II (500 °C). In this

Figure 3. TGA and DTG curves of [Ca−Fe−Cl] LDH in (a) flowing


N2 and (b) flowing CO2.
Figure 5. TGA and DTG curves of [Ca−Al−NO3] LDH in flowing
CO2.
This mass loss is realized over three steps at 161 °C
(14.4%), 278 °C (12.1%), and 672 °C (9.4%). In flowing CO2,
a behavior similar to that seen in [Ca−Al−Cl] is observed. At case, the decomposition is incomplete at 900 °C (observed
the end of the second step of the mass loss (26.1%, 400 °C, mass loss 30%) and correspondingly, there is no mass gain in
plateau-I), the dehydroxylated residue is X-ray amorphous the cooling cycle. The residue obtained at 900 °C was
(Figure 4). This residue does not gain mass on cooling, but on dissolved in dilute HCl and analyzed by ion chromatography.
heating in CO2, a mass gain corresponding to 8.5% of the Residual NO3− ion was observed to the extent of 0.07 mol %
original LDH is observed at plateau-II (680 °C) to yield (4.5 wt %) with respect to LDH and 0.10 mol % (6.3 wt %)
CaCO3. The absence of any CaCO3 at plateau-I is responsible with respect to residue, showing that the decomposition of the
for the higher uptake when compared to the [Ca−Al] LDH. LDH is incomplete under flowing CO2 and the nitrate is
This mass gain corresponds to the uptake of 1.93 mmol g−1 of retained in the residue. The nitrate-containing phase could not
the original LDH. Above 700 °C, there is a steep mass loss be identified. The phase obtained isothermally at plateau-I was
(18.3%) until 900 °C because of the decomposition of the a mixture of CaCO3 and Ca(OH)2 (Figure S6). The phase
3201 DOI: 10.1021/acsomega.9b00083
ACS Omega 2019, 4, 3198−3204
ACS Omega Article

obtained at plateau-II was CaCO3. At plateau-III, a mix of


CaCO3, mayenite, and another ternary oxide Ca5(Al3O7)2 was
observed.

3. DISCUSSION
The LDH crystallizes in a rhombohedral structure. CaO
crystallizes in the cubic rock salt structure. There is a well-
established topotactic relationship between the precursor and
product phases, which can be described as [00l]H/R∥[111]C (l
= 1 for H: hexagonal; l = 3 for R: rhombohedral; C: cubic).25
The decomposition of the LDH takes place by the
condensation of the metal hydroxide layers brought about by
dehydration and dehydroxylation. This occurs just below 400
°C. CaO is therefore expected to form throughout the bulk of
the LDH. The uptake of CO2 takes place thereafter at T > 460
Figure 6. TGA and DTG curves of CaO in flowing CO2.
°C. The uptake of CO2 is an addition reaction of the type CaO
+ CO2 → CaCO3. However, all of the Ca are not utilized in
the first cycle. If indeed all of the free CaO were to be utilized, shows no signs of decomposition at 900 °C. In the case of
plateau-II would correspond to 86% of the mass of the [Ca− [Ca−Fe−Cl] LDH, the formed CaCO3 decomposes at 800 °C
Fe] LDH (91% in the case of [Ca−Al] LDH). The observed itself bringing about a 100 °C reduction in the thermal penalty
mass gain at plateau-II corresponds to ∼26% Ca utilization compared to the [Ca−Al−Cl] LDH. The combined/
(∼11% in the case of the [Ca−Al] LDH). CaCO3 formation synergistic effect of Fe and Cl makes the [Ca−Fe−Cl] LDH
evidently takes place on the exposed surface of the CaO interesting for practical use. In all cases isothermal treatment of
crystallites. Because continued stay (up to 2 h) at various the LDH at temperatures of the three plateaus did not enhance
temperatures spanned by plateau-II did not lead to further the mass gain, showing that the mass gain in the TGA is not
mass gain (Figure S3), it is our conclusion that CaO utilization limited by kinetics.
is not limited by the kinetics of CO2 uptake, but by the uptake
of CO2 at the surface of the CaO crystallites. The CaCO3 4. CONCLUSIONS
formed on the surface prevents a more complete material Hydrocalumite-like LDHs are superior to the hydrotalcite-like
utilization. compounds for applications related to mineralization of CO2.
The residue obtained at plateau-II was calcined in flowing These Ca-containing phases are single-source precursors for
nitrogen at 900 °C and reheated in flowing CO2. The oxide both CaO and the ternary oxide supports. Among the
residue gained nearly 30% mass by aggressive uptake of CO2 in hydrocalumites, the [Ca−Fe−Cl] phase shows the highest
the temperature range of 434−685 °C (Figure S7). We uptake of CO2. The product of CO2 uptake is the calcite form
conclude that intermediate calcination at high temperature of CaCO3. The presence of intercalated chloride ions appears
exposes fresh CaO surface, facilitating increased material to reduce the thermal penalty when compared to CaO,
utilization. This is in direct contrast to silica removal by whereas intercalated nitrate ions fail to induce the decom-
hydrotalcite-like LDHs.26,27 Silicates being nonvolatile ions position of the CaCO3 formed as a result of CO2 uptake.
accumulate in the solid phase on repeated cycling, leading to
decreased efficiency. Carbonates are volatile ions and behave 5. EXPERIMENTAL SECTION
differently. The [Ca−M−X] LDHs (M = Al and Fe; X = Cl− and NO3−)
Pure CaO is used for high-temperature CO2 cycling. were synthesized by coprecipitation. In separate experiments,
However, there are two major drawbacks of CaO: (i) the the [Ca−Al−Cl] and [Ca−Fe−Cl] LDHs were prepared by
thermal penalty is high.28,29 The CO2 uptake temperature is dropwise addition of a mixed metal chloride solution in the
close to 650 °C for CaO and 450 °C for Ca(OH)2, whereas no stoichiometric ratio ([Ca]/[M] = 2, dropping rate 0.3 mL
decomposition of the formed CaCO3 is observed up to 900 °C min−1) to a NaCl solution (100 mL; 0.4 mol L−1) containing
(Figure 6) in both cases. (ii) The CO2 uptake falls sharply by five times excess Cl− ions. The pH of 11.5 was maintained for
80% during thermal cycling;30 the fall is attributed to high- all of the syntheses by dispensing NaOH solution (0.5 mol
temperature sintering of the CaO crystallites.30,31 L−1) at 60 °C using a Metrohm model 836 Titrando, operating
There have been major technological and scientific efforts to in the pH STAT mode. N2 gas was bubbled during
alleviate these problems by (i) appropriate engineering to precipitation and ageing to avoid carbonate contamination.
utilize the heat generated by the exothermicity of the Once the precipitation was complete, the samples were aged in
carbonation cycle and reduce the thermal penalty32,33 and mother liquor under a N2 blanket for 18 h under vigorous
(ii) by dispersing CaO on an inert oxide support such as stirring.
mineral mayenite to prevent agglomeration and sintering of the In the case of [Ca−Al−NO3] LDH, rapid coprecipitation
active CaO crystallites.34,35 was carried out by mixing the mixed metal nitrate solution and
The use of hydrocalumite-based Ca-rich LDH as a precursor NaOH (50 mL, 2 mol L−1) in single step in a screw cap bottle.
has the potential to generate in single step, active CaO, as well NaNO3 (15 times in excess) was added to the NaOH solution
as mayenite (Ca2Fe2O5 in the [Ca−Fe] LDH) for effective prior to mixing. The pH of the final mixture was >12. The
CO2 uptake. Further, the use of the intercalated Cl− ion coprecipitated slurries were aged in mother liquor (48 h, T 85
induces the decomposition of the formed CaCO3 at 900 °C in °C).
the case of [Ca−Al−Cl] LDH. In contrast with this, the All of the precipitates were centrifuged, washed with boiled
CaCO3 formed by the uptake of CO2 by pure CaO (Figure 6) type II water (specific resistance 15 MΩ cm, Millipore
3202 DOI: 10.1021/acsomega.9b00083
ACS Omega 2019, 4, 3198−3204
ACS Omega


Article

Academic water purification system), and dried in a hot air REFERENCES


oven at 60 °C. (1) Pacala, S.; Socolow, R. Stabilization Wedges: Solving the Climate
5.1. Wet Chemical Analysis. The compositions of the Problem for the Next 50 Years with Current Technologies. Science
LDH were determined by a combination of techniques. The 2004, 305, 968−972.
Ca and Fe contents were determined by atomic absorption (2) Roth, E. A.; Agarwal, S.; Gupta, R. K. Nanoclay-Based Solid
spectroscopy using a Shimadzu model AA-6650 atomic Sorbents for CO2 Capture. Energy Fuels 2013, 27, 4129−4136.
absorption spectrometer. The Al content was estimated (3) Aresta, M.; Dibenedetto, A. Utilisation of CO2 as a chemical
gravimetrically by dissolving a preweighed amount of the feedstock: opportunities and challenges. Dalton Trans. 2007, 2975−
LDH in a mineral acid and precipitating Al3+ in an ammoniacal 2992.
buffer. The resulting gelatinous precipitate was calcined (T ≈ (4) Langanke, J.; Wolf, A.; Hofmann, J.; Böhm, K.; Subhani, M. A.;
900 °C) and weighed as Al2O3. The anion contents were Müller, T. E.; Leitner, W.; Gürtler, C. Carbon dioxide (CO2) as
sustainable feedstock for polyurethane production. Green Chem. 2014,
determined by dissolving an accurately weighed amount of the
16, 1865−1870.
LDH in a suitable mineral acid. The anions released were (5) Boot-Handford, M. E.; Abanades, J. C.; Anthony, E. J.; Blunt, M.
estimated by ion chromatography (Metrohm model 861 J.; Brandani, S.; Mac Dowell, N.; Fernández, J. R.; Ferrari, M.-C.;
advanced compact ion chromatograph with Metrosep SUP5 Gross, R.; Hallett, J. P.; et al. Carbon Capture and Storage Update.
150 column). Energy Environ. Sci. 2014, 7, 130−189.
5.2. Characterization. All samples were characterized by (6) Dean, C. C.; Blamey, J.; Florin, N. H.; Al-Jeboori, M. J.; Fennell,
PXRD using a Bruker D8 ADVANCE powder diffractometer P. S. The calcium looping cycle for CO2 capture from power
(source Cu Kα radiation, Ni filter, λ = 1.5418 Å) operating in generation, cement manufacture and hydrogen production. Chem.
reflection geometry. The PXRD patterns of the LDHs were Eng. Res. Des. 2011, 89, 836−855.
indexed by using APPLEMAN program, part of the code (7) Wang, J.; Huang, L.; Yang, R.; Zhang, Z.; Wu, J.; Gao, Y.; Wang,
PROSZKI suite of programs. The PXRD patterns of the oxide Q.; O’Hare, D.; Zhong, Z. Recent advances in solid sorbents for CO2
residues were indexed by “finger printing” by comparing the capture and new development trends. Energy Environ. Sci. 2014, 7,
observed 2θ values with those reported in the powder 3478−3518.
(8) Rives, V. Characterisation of Layered Double Hydroxides and
diffraction files. TGAs were carried out using a Mettler Toledo Their Decomposition Products. Mater. Chem. Phys. 2002, 75, 19−25.
TGA/SDTA 851e system (25−900 °C, heating rate 5 °C (9) Radha, S.; Navrotsky, A. Energetics of CO2 Adsorption on Mg-
min−1) driven by STARe 7.01 software. The analyses were Al Layered Double Hydroxides and Related Mixed Metal Oxides. J.
carried out in flowing N2 (control) and flowing CO2 (test) to Phys. Chem. C 2014, 118, 29836−29844.
measure the mass gain if any, as a result of CO2 uptake. In (10) Yong, Z.; Rodrigues, A. E. Hydrotalcite-like Compounds as
separate experiments, the LDHs were heated in the TGA Adsorbents for Carbon Dioxide. Energy Convers. Manage. 2002, 43,
balance at different temperatures determined by the appear- 1865−1876.
ance of plateaus in the TGA profile. The phases obtained were (11) Wang, Q.; Wu, Z.; Tay, H. H.; Chen, L.; Liu, Y.; Chang, J.;
analyzed by PXRD. TGA measurements were also carried out Zhong, Z.; Luo, J.; Borgna, A. High temperature adsorption of CO2
on CaO and Ca(OH)2 used as controls. IR spectra were on Mg-Al hydrotalcite: Effect of the charge compensating anions and
recorded by using a Bruker model Alpha-P IR spectrometer the synthesis pH. Catal. Today 2011, 164, 198−203.
(12) Sun, Q.; Li, Z.; Searles, D. J.; Chen, Y.; Lu, G.; Du, A. Charge-
(Diamond ATR cell, 400−4000 cm−1, 4 cm−1 resolution).


Controlled Switchable CO2 Capture on Boron Nitride Nanomaterials.
J. Am. Chem. Soc. 2013, 135, 8246−8253.
ASSOCIATED CONTENT (13) Lee, J. M.; Min, Y. J.; Lee, K. B.; Jeon, S. G.; Na, J. G.; Ryu, H.
* Supporting Information
S J. Enhancement of CO2 Sorption Uptake on Hydrotalcite by
The Supporting Information is available free of charge on the Impregnation with K2CO3. Langmuir 2010, 26, 18788−18797.
ACS Publications website at DOI: 10.1021/acsome- (14) Constantino, V. R. L.; Pinnavaia, T. J. Basic Properties of
ga.9b00083. Mg2+1‑xAl3+x Layered Double Hydroxides Intercalated by Carbonate,
Hydroxide, Chloride, and Sulfate Anions. Inorg. Chem. 1995, 34,
PXRD patterns of the pristine hydrocalumite-like LDHs; 883−892.
tables of 2θ values and indexing of the reflections of the (15) Kumar, P. P.; Kalinichev, A. G.; Kirkpatrick, R. J. Hydration,
hydrocalumite-like LDHs; infrared spectra and TGA Swelling, Interlayer Structure, and Hydrogen Bonding in Organo-
profiles of phases obtained at the different plateau layered Double Hydroxides: Insights from Molecular Dynamics
temperatures; and TGA data of the nitrate-intercalated Simulation of Citrate-Intercalated Hydrotalcite. J. Phys. Chem. B
LDH (PDF) 2006, 110, 3841−3844.


(16) Marappa, S.; Kamath, P. V. Interaction of pristine hydrotalcite-
like layered double hydroxides with CO2: a thermogravimetric study.
AUTHOR INFORMATION Bull. Mater. Sci. 2015, 38, 1783−1790.
Corresponding Author (17) Vieille, L.; Rousselot, I.; Leroux, F.; Besse, J.-P.; Taviot-Guého,
*E-mail: vishnukamath8@hotmail.com. C. Hydrocalumite and Its Polymer Derivatives. 1. Reversible Thermal
Behavior of Friedel’s Salt: A Direct Observation by Means of High-
ORCID Temperature in Situ Powder X-ray Diffraction. Chem. Mater. 2003,
P. Vishnu Kamath: 0000-0002-3549-7024 15, 4361−4368.
Notes (18) Radha, A. V.; Forbes, T. Z.; Killian, C. E.; Gilbert, P. U. P. A.;
The authors declare no competing financial interest. Navrotsky, A. Transformation and Crystallization Energetics of


Synthetic and Biogenic Amorphous Calcium Carbonate. Proc. Natl.
ACKNOWLEDGMENTS Acad. Sci. U.S.A. 2010, 107, 16438−16443.
(19) Rajamathi, M.; Nataraja, G. D.; Ananthamurthy, S.; Kamath, P.
Authors thank the Department of Science and Technology, V. Reversible Thermal Behavior of the Layered Double Hydroxide of
Government of India, for financial support. Authors also thank Mg with Al: Mechanistic Studies. J. Mater. Chem. 2000, 10, 2754−
the Reviewers for their useful comments. 2753.

3203 DOI: 10.1021/acsomega.9b00083


ACS Omega 2019, 4, 3198−3204
ACS Omega Article

(20) Prasanna, S. V.; Vishnu Kamath, P. Chromate Uptake


Characteristics of the Pristine Layered Double Hydroxides of Mg
with Al. Solid State Sci. 2008, 10, 260−266.
(21) Radha, A. V.; Vishnu Kamath, P.; Shivakumara, C. Mechanism
of the Anion Exchange Reactions of the Layered Double Hydroxides
(LDHs) of Ca and Mg with Al. Solid State Sci. 2005, 7, 1180−1187.
(22) Rousselot, I.; Taviot-Guého, C.; Leroux, F.; Léone, P.;
Palvadeau, P.; Besse, J.-P. Insights on the Structural Chemistry of
Hydrocalumite and Hydrotalcite-like Materials: Investigation of the
Series Ca2M3+(OH)6Cl·2H2O (M3+: Al3+, Ga3+, Fe3+, and Sc3+) by X-
Ray Powder Diffraction. J. Solid State Chem. 2002, 167, 137−144.
(23) Allada, R. K.; Pless, J. D.; Nenoff, T. M.; Navrotsky, A.
Thermochemistry of Hydrotalcite-like Phases Intercalated with CO32‑,
NO3‑, Cl‑, I‑, and ReO4‑. Chem. Mater. 2005, 17, 2455−2459.
(24) Bontchev, R. P.; Liu, S.; Krumhansl, J. L.; Voigt, J.; Nenoff, T.
M. Synthesis, Characterization, and Ion Exchange Properties of
Hydrotalcite Mg6Al2(OH)16(A)x(A′)2‑x·4H2O (A, A′ = Cl‑, Br‑, I‑, and
NO3‑, 2 ≥x≥ 0) Derivatives. Chem. Mater. 2003, 15, 3669−3675.
(25) Figlarz, M.; Gérand, B.; Delahaye-Vidal, A.; Dumont, B.; Harb,
F.; Coucou, A.; Fievet, F. Topotaxy, nucleation and growth. Solid
State Ionics 1990, 43, 143−170.
(26) Sasan, K.; Brady, P. V.; Krumhansl, J. L.; Nenoff, T. M.
Removal of dissolved silica from industrial waters using inorganic ion
exchangers. J. Water Process Eng. 2017, 17, 117−123.
(27) Sasan, K.; Brady, P. V.; Krumhansl, J. L.; Nenoff, T. M.
Exceptional selectivity for dissolved silicas in industrial waters using
mixed oxides. J. Water Process Eng. 2017, 20, 187−192.
(28) Arias, B.; Alonso, M.; Abanades, C. CO2 Capture by Calcium
Looping at Relevant Conditions for Cement Plants: Experimental
Testing in a 30 kWth Pilot Plant. Ind. Eng. Chem. Res. 2017, 56,
2634−2640.
(29) Dunstan, M. T.; Maugeri, S. A.; Liu, W.; Tucker, M. G.; Taiwo,
O. O.; Gonzalez, B.; Allan, P. K.; Gaultois, M. W.; Shearing, P. R.;
Keen, D. A.; et al. In situ studies of materials for high temperature
CO2 capture and storage. Faraday Discuss. 2016, 192, 217−240.
(30) Fennell, P. S.; Pacciani, R.; Dennis, J. S.; Davidson, J. F.;
Hayhurst, A. N. The Effects of Repeated Cycles of Calcination and
Carbonation on a Variety of Different Limestones, as Measured in a
Hot Fluidized Bed of Sand. Energy Fuels 2007, 21, 2072−2081.
(31) Anthony, E. J. Solid Looping Cycles: A New Technology for
Coal Conversion. Ind. Eng. Chem. Res. 2008, 47, 1747−1754.
(32) Li, Z.-s.; Cai, N.-s.; Huang, Y.-y. Effect of Preparation
Temperature on Cyclic CO2 Capture and Multiple Carbonation−
Calcination Cycles for a New Ca-Based CO2 Sorbent. Ind. Eng. Chem.
Res. 2006, 45, 1911−1917.
(33) Kierzkowska, A. M.; Pacciani, R.; Müller, C. R. CaO-Based CO2
Sorbents: From Fundamentals to the Development of New, Highly
Effective Materials. ChemSusChem 2013, 6, 1130−1148.
(34) Dennis, J. S.; Pacciani, R. The rate and extent of uptake of CO2
by a synthetic, CaO-containing sorbent. Chem. Eng. Sci. 2009, 64,
2147−2157.
(35) Silaban, A.; Narcida, M.; Harrison, D. P. Characteristics of the
reversible reaction between CO2(g) and calcined dolomite. Chem.
Eng. Commun. 1996, 146, 149−162.

3204 DOI: 10.1021/acsomega.9b00083


ACS Omega 2019, 4, 3198−3204

You might also like