Subscriber access provided by UniSA Library
Energy and the Environment
Accurate Control of Cage-Like CaO Hollow Microspheres for Enhanced CO2 Capture in Calcium Looping via a Template-assisted Synthesis Approach Jian Chen, Lunbo Duan, and Zhao Sun Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.8b06138 • Publication Date (Web): 18 Jan 2019 Downloaded from http://pubs.acs.org on January 18, 2019
Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.
is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.
Page 1 of 32
Environmental Science & Technology
1
Accurate Control of Cage-Like CaO Hollow
2
Microspheres for Enhanced CO2 Capture in Calcium
3
Looping via a Template-assisted Synthesis
4
Approach
5
Jian Chen, Lunbo Duan*, Zhao Sun
6
Key Laboratory of Energy Thermal Conversion and Control, Ministry of Education, School of
7
Energy and Environment, Southeast University, Nanjing 210096, China
8
ABSTRACT: Herein we report the development of synthetic CaO-based sorbents for enhanced
9
CO2 capture in calcium looping via a template-assisted synthesis approach, where carbonaceous
10
spheres (CSs) derived from hydrothermal reaction of starch are used as the templates. Cage-like
11
CaO hollow microspheres are successfully synthesized only using urea as the precipitant, and the
12
formation mechanism of this unique hollow microsphere structure is discussed deeply. Moreover,
13
cage-like CaO hollow microspheres possess an initial carbonation conversion of 98.2% and 82.5%
14
under a mild and harsh conditions, respectively. After the fifteen cycles, cage-like CaO hollow
15
microspheres still possess a carbonation value of 49.2% and 39.7% under the corresponding
16
conditions, exceeding the reference limestone by 85.7% and 148.1%, respectively. Two kinetic
17
models are used to explore the mechanism of carbonation reaction for cage-like CaO hollow
ACS Paragon Plus Environment
1
Environmental Science & Technology
Page 2 of 32
18
microspheres, which are subsequently proved to be feasible for analysis of chemical-controlled
19
stage and diffusion-controlled stage in the carbonation process. It is found the unique hollow
20
microsphere structure can significantly reduce the activation energy of carbonation reaction
21
according to the kinetic calculation. Furthermore, the energy and raw material consumptions
22
related to the synthesis of cage-like CaO hollow microspheres are analyzed by the life cycle
23
assessment (LCA) method.
24 25 26
INTRODUCTION
27
Calcium looping, based on a reversible carbonation/calcination reaction of CaO-based
28
sorbents, is recognized as one of the most promising CO2 capture technologies.1-12 However, the
29
CO2 uptake capacity of naturally occurring CaO-based sorbents decreases significantly with the
30
repeated carbonation/calcination cycles.13,14 Thus, to stabilize the CO2 capture performance of
31
CaO-based sorbents, different strategies such as hydration13,15 and thermal pre-treatment16 have
32
been proposed. Unfortunately, these methods can only partially mitigate the dramatic decline in
ACS Paragon Plus Environment
2
Page 3 of 32
Environmental Science & Technology
33
the CO2 uptake of CaO-based sorbents.17,18 The development of synthetic CaO-based sorbents has
34
received considerable attention recently.19 It has been reported CaO-based sorbents synthesized
35
with a nanostructure possessed a high and stable CO2 uptake capacity, which can minimize the
36
diffusion lengths of CO2 through the freshly formed CaCO3 product layer.20-22 Furthermore, the
37
CaO-based sorbents should possess a certain level of porosity in order to be able to compensate
38
for the pronounced volume variations during the repeated cycles, as the molar volume of CaCO3
39
is more than twice as high as that of CaO.23
40
Hollow structures refer to materials with well-defined boundaries and interior cavities, which
41
have attracted tremendous attention due to their unique structure and controllable morphology.24-
42
26
43
assisted synthesis approach, using different templates such as carbonaceous spheres (CSs),27
44
polystyrene microbeads,28 sulfonated polystyrene,29 and carbon gel.30 Recently, CaO hollow
45
microspheres have been proposed for calcium looping process due to their significant potential to
46
enhance CO2 capture performance. On one hand, a hollow microsphere structure can provide a
47
high specific surface area with numerous active sites for the occurrence of carbonation reaction,
48
enabling rapid kinetics and high sorption capacity. On the other hand, the central void can provide
49
the required volume to buffer against the pronounced volume variations accompanied with the
50
repeated carbonation/calcination cycles, thus leading to improved cycling stability. Ping et al.
51
synthesized CaO hollow microspheres using CSs as the template, and found that CaO hollow
52
microspheres with a cavity diameter of 1.62 μm reached the maximal theoretical value (i.e., 0.786
53
g CO2/g CaO).27 Noted that CS templates are more environmental-friendly compared to the
54
polymer sphere templates such as polystyrene microbeads28 and sulfonated polystyrene.29 It is
55
because the synthetic procedure of CS templates involves none of the toxic organic solvents,
The most widely employed method for the production of hollow structures is the template-
ACS Paragon Plus Environment
3
Environmental Science & Technology
Page 4 of 32
56
initiators or surfactants, which are commonly used in the fabrication of polymer sphere
57
templates.31,32 However, the work conducted by Ping et al.27 studied the CO2 sorption capacity and
58
sorption rate of CaO hollow microspheres only in the first cycle, which is lack of research for
59
cyclic stability. And their experiments were conducted under an unrealistic operating condition,
60
i.e., regeneration in pure N2 atmosphere at a low temperature of 800 °C. Besides, the essential
61
conditions of fabrication of CaO hollow microspheres were not clearly obtained in their work.
62
Although CaO hollow microspheres exhibit great potential in enhancement of CO2 capture
63
performance, yet little work has been conducted concerning the fabrication of CaO hollow
64
microspheres. Thus, here we use the template-assisted synthesis approach to yield a highly
65
effective CaO-based sorbent. CSs, formed from a hydrothermal reaction of soluble starch, are used
66
as the templates. The influence of different precipitants, Ca2+ concentration, and the size of CS
67
templates on the morphologies of the as-synthesized CaO-based sorbents are systematically
68
investigated. Cage-like CaO hollow microspheres featuring highly porous shells can be accurately
69
synthesized, and the formation mechanism of this unique hollow microsphere structure is
70
discussed deeply. Moreover, the cyclic CO2 capture performance of the synthetic CaO-based
71
sorbents is assessed critically under both mild and harsh conditions, which was also compared to
72
the reference limestone. Kinetic study using two different models is conducted to understand the
73
reaction mechanism at different stages in the carbonation process of the cage-like CaO hollow
74
microspheres. Furthermore, the energy and raw material consumptions related to the synthesis of
75
cage-like CaO hollow microspheres are analyzed by the life cycle assessment (LCA) method.
76
EXPERIMENTAL SECTION
ACS Paragon Plus Environment
4
Page 5 of 32
Environmental Science & Technology
77
For simplicity, here the abbreviations of CaO-U, CaO-SC, and CaO-OA are used to denote
78
CaO-based sorbents prepared using urea, sodium carbonate and oxalic acid as the precipitant,
79
respectively. The detailed information concerning preparation of CS templates, synthesis of CaO-
80
based sorbents, characterization and performance test can be found in Supporting Information (SI).
81
RESULTS AND DISCUSSION
82
Characterizations of CS templates. CS templates are fabricated via the hydrothermal
83
reaction of soluble starch. It is reported by Sun et al. that the diameters of CSs were influenced by
84
hydrothermal temperature, duration and concentration of the starting material,32 which would, in
85
turn, affect the morphologies of the resultant materials.27,33 Therefore, we systematically
86
investigate the influence of synthesis parameters, including hydrothermal temperature, duration
87
and concentration of starch, on the morphologies and sizes of CS templates, aiming to synthesize
88
monodispersed ones with controllable sizes.
89
As shown in Figure S1 (SI), the starting material, i.e., soluble starch, exhibits very irregular
90
morphologies. After hydrothermal treatment at 180 °C with a duration of 3 h, CSs with smooth
91
surfaces and narrow particle size distributions can be obtained. However, for a short hydrothermal
92
duration of 1.5 h, no CSs form; whereas CSs with non-uniform particles size can be observed for
93
the prolonged durations, i.e., over 3 h. Moreover, a hydrothermal temperature below or over 180
94
°C will lead to coarse surface or wide particle size distribution (Figure S2). It is worth mentioning
95
that no CSs form when the hydrothermal temperature is as low as 140 °C. Similar results can be
96
found in other literature.32,34,35 Furthermore, it seems that soluble starch concentration plays an
97
important role in particle size distributions of CS templates. The particle sizes of templates increase
98
with the increase of starch concentrations, as given in Figure 1 and Table 1. Thus, we can
ACS Paragon Plus Environment
5
Environmental Science & Technology
Page 6 of 32
99
accurately control the size of CS templates by adjusting the soluble starch concentrations under a
100
hydrothermal temperature of 180 °C with a duration of 3 h. For simplicity, the CS templates
101
derived from starch concentrations of 50, 100, and 200 g starch/L·H2O are denoted as T-1, T-2,
102
and T-3, respectively.
103
The XRD pattern of T-3 (Figure 2a) exhibits only a broad peak between 20° and 30°, which
104
corresponds to indicative of amorphous carbon and is a characteristic of colloidal CSs.36-38
105
Moreover, the adsorption band present in Figure 2b at 1709 and 1613 cm−1 can be assigned to C=O
106
and C=C vibrations, respectively, suggesting the dehydrative aromatization of starch during
107
hydrothermal treatment.32 The bands observed at 3400 and 1000-1300 cm-1 are assigned to
108
stretching and bending vibrations of C=O and OH groups, implying the existence of large numbers
109
of residual hydroxy groups.39 These results indicate that the surface of CS templates is hydrophilic
110
and enriched with OH, C=O, C=C groups, which will be able to bind metal cations through
111
coordination or electrostatic interactions.32,40,41 Furthermore, the Raman spectra of T-3 (Figure S3)
112
shows two broad overlapping bands at around 1360 and 1587 cm-1, which is typical of carbonized
113
material.42 The G band (1587 cm-1) is assigned to the in-plane bond-stretching motion of pairs of
114
C sp2 atoms both in aromatic and olefinic molecules,43 whereas the D band (1360 cm-1) is assigned
115
to ring-breathing vibrations in benzene or condensed benzene rings in amorphous (partially
116
hydrogenated) carbon films.44 Therefore these results confirm the existence of small aromatic
117
clusters in T-3, which agrees well with the aromatization of the materials observed by FTIR
118
spectra.
ACS Paragon Plus Environment
6
Page 7 of 32
Environmental Science & Technology
119 120
Figure 1. SEM images and particle size distributions of the CS templates prepared under
121
different concentration of starch with a hydrothermal temperature of 180 °C and a duration of 3
122
h. (a), (d): T-1; (b), (e): T-2; (c), (f): T-3.
123
Table 1. Particle size, zeta potential, and pH value of different CS templates Sample
T-1
T-2
T-3
Mean particle size (μm)
0.44
0.84
2.19
Zeta potential (mV)
-16.0
-21.5
-33.0
pH value
7.2
7.3
7.3
124
ACS Paragon Plus Environment
7
Environmental Science & Technology
Page 8 of 32
125 126
Figure 2. Characterizations of CS templates and CaO-based sorbents. (a): Typical XRD patterns
127
of sample a’: T-3, sample b’: T-3 after Ca2+ ions deposition in the presence of urea, and sample
128
c’: sample b’ after template removal by calcination. (b): Typical FTIR spectra of T-3 before and
129
after Ca2+ ions deposition.
130
Formation mechanism of synthetic CaO-based sorbents with different structures. Zeta
131
potential analysis on the CS templates indicates that the surfaces of the templates are negatively
132
charged in ethanol (given in Table 1). However, this value is changed to -3.65 mV (pH=8.0) after
133
Ca2+ ions deposition on T-3 with the addition of urea. Moreover, as clearly shown in Figure S4,
134
the surface of T-3 after Ca2+ ions deposition becomes much coarser, which is coated by a product
135
layer. On the basis of the XRD result for T-3 after Ca2+ ions deposition (Figure 2a), except for the
136
broad peak corresponding to indicative of amorphous carbon, CaCO3 is also detected. The product
137
layer is thus identified as CaCO3, and the increase of zeta potential value after Ca2+ deposition
138
results from the Ca2+ ions deposition on the surface of T-3, i.e., CaCO3 product layer.
139
Consequently, the increase of zeta potential value after Ca2+ ions deposition in this work suggests
140
Ca2+ ions are deposited into the surface layer of the templates. For the case of CS template after
141
Ca2+ ions deposition, FTIR results reveal that the absorption bands weaken in general, especially
ACS Paragon Plus Environment
8
Page 9 of 32
Environmental Science & Technology
142
for the OH group at 3400 cm-1 (Figure 2b). This result keeps consistent with the work conducted
143
by Armutlulu et al.39 and Sun et al..32 After template removal, only CaO is detected, suggesting
144
the complete removal of the CS templates by calcination (Figure 2a).27
145
Figure 3 gives the surface morphologies of CaO-based sorbents prepared using different
146
precipitants, including sodium carbonate, oxalic acid, and urea. It is interesting to find that the
147
precipitant exerts an important role in the morphologies of synthetic CaO-based sorbents. When
148
sodium carbonate or oxalic acid is used as the precipitant, loose structure, rather than hollow one,
149
can be obtained (Figure 3a-d). However, cage-like hollow microsphere structures featuring highly
150
porous shells can be obtained with the addition of urea (Figure 3e, f). The presence of highly
151
porous shells is critical due to it allows a rapid transport of CO2 to and from the material. Moreover,
152
the porous shell is consisted of CaO nanoparticles (Figure S5), thus satisfying an essential
153
requirement to avoid (or at least minimize) diffusion limitations during the CO2 capture reaction.23
154
Figure 3g shows a typical TEM image of the cage-like CaO hollow microspheres, further
155
confirmed the hollow interiors clearly by the obvious electron-density difference between the dark
156
edge and the pale center. And the cage-like CaO hollow microspheres are polycrystalline
157
according to the diffraction pattern (Figure 3h).
158
ACS Paragon Plus Environment
9
Environmental Science & Technology
Page 10 of 32
159
Figure 3. SEM images of (a), (b): CaO-SC, (c), (d): CaO-OA, and (e), (f): CaO-U; TEM image
160
(g) and SAED pattern (h) of CaO-U. Each CaO-based sorbent was prepared from 0.5 M
161
Ca(NO3)2 solution and used T-3 as the template.
162
The entirely different morphologies of synthetic CaO-based sorbents prepared using different
163
precipitants can be probably ascribed to both the formation route and the amount of the precipitated
164
CaCO3. When sodium carbonate or oxalic acid is added into the Ca(NO3)2 solution, the vast
165
majority of Ca2+ ions are precipitated as CaCO3 or CaC2O4 immediately, which subsequently
166
agglomerate and wrap around the templates randomly, as shown in Figure 4B. Hence, CaO-based
167
sorbent with loose structure can be obtained after removal of templates. Despite hollow
168
microsphere structure is not obtained with the addition of sodium carbonate or oxalic acid, the
169
removal of CS templates results in an enhancement of porosity, which is beneficial to CO2 capture.
170
Duan et al.45 synthesized CaO-based sorbent using commercial flour as the biomass-template, and
171
found that the CO2 capture capacity was significantly enhanced due to the improved porosity with
172
high specific surface area and pore volume. Similar conclusion was also drawn by Ridha et al.,46
173
who prepared four types of biomass-templated CaO-based sorbents and found that biomass
174
materials improved the porosity of sorbents. Whereas with the addition of urea at room
175
temperature, a clean solution can be obtained without any precipitated CaCO3. This is due to the
176
+ 2gradually hydrolyses of urea, i.e., CO NH 2 2 +2H 2 O 2NH 4 CO3 , occurs only at a little
177
higher temperature (>75 °C).27,33,39 Once the solution is transferred into the water bath (90 °C),
178
Ca2+ ions are precipitated gradually as CaCO3 with the progressive hydrolysis of urea. At the same
179
time, the amount of the precipitated Ca2+ ions using urea as the precipitant is significantly reduced
180
compared to the conditions where sodium carbonate or oxalic acid is used as the precipitant, mainly
181
because of the incomplete hydrolysis of urea.47 Both of them are beneficial to the uniform coating
ACS Paragon Plus Environment
10
Page 11 of 32
Environmental Science & Technology
182
of CaCO3 around the templates, as shown in Figure 4C. It should be noted that this synthesis
183
approach is distinctively different from the surface adsorption method of metal cations,48 since the
184
precipitation process is not limited by the adsorption capacity of the surface (compare to Figure
185
4A and Figure 4B, C). Finally, the removal of carbon cores, and densification, cross-link, and
186
phase transformation of CaCO3 in the surface layer via calcination lead to the formation of cage-
187
like CaO hollow microspheres.
188 189
Figure 4. Schematic illustrations of the formation mechanism of CaO-based sorbents
190
synthesized (A): without any precipitants, with the addition of (B): sodium carbonate or oxalic
191
acid and (C): urea.
192
Further investigation demonstrates that the formation of cage-like CaO hollow microspheres
193
is strongly dependent on experimental parameters such as Ca(NO3)2 concentration, the presence
194
of template, the particle size of template, the addition of urea, and the heating rate during template
195
removal process. When the Ca(NO3)2 concentration is reduced to 0.2 M, cage-like CaO hollow
ACS Paragon Plus Environment
11
Environmental Science & Technology
Page 12 of 32
196
microspheres with a smaller particle size can be obtained, as shown in Figure 5a, b. This means
197
we can accurately control the particle size of CaO hollow microspheres via the adjusting of
198
Ca(NO3)2 concentration. Similar results were also obtained in the preparations of SnO241 and
199
Fe2O3.33 Moreover, it seems that the addition of both urea and template are essential to obtain cage-
200
like CaO hollow microspheres. In the absence of urea, loose structure rather than cage-like hollow
201
microsphere one can be obtained (Figure 5c). This can be explained by the fact that the amount of
202
Ca2+ ions deposited on the surface of templates is significantly decreased, mainly due to the limit
203
of the adsorption capability of Ca2+ ions by the hydrophilic surface within such a short deposition
204
duration (i.e., 4 h). As a result, the shell walls of hollow microspheres are so thin, which easily
205
collapsed during template removal process.33 This result also further proves that the synthesis
206
method using precipitants is the precipitation strategy instead of the surface adsorption route. In
207
the absence of template (Figure 5d, e), it is interesting to find that CaO-based sorbent exhibits long
208
strip shapes rather than cage-like hollow microspheres. Besides, a worse porosity is observed
209
compared to the sorbents prepared with the addition of template, mainly due to the lack of gaseous
210
product release during the template removal process. It thus also implies that the addition of
211
template is beneficial to CO2 capture, whether cage-like hollow microsphere structures are formed
212
finally.
213
The heating rate during the template removal process also plays an important role in the
214
formation of cage-like CaO hollow microspheres. As shown in Figure 5f, when the template
215
removal process is performed directly at 800 °C, there are many fragments and broken cage-like
216
CaO hollow microspheres, rather than complete microspheres. It results from the significant
217
thermal stress and intense gas release caused by the fast calcination of templates. Furthermore, it
218
is interesting to find that when T-2 or T-1 is used as the template, i.e., a reduction of template
ACS Paragon Plus Environment
12
Page 13 of 32
Environmental Science & Technology
219
particle size from 2.19 μm to 0.84 μm or 0.44 μm, respectively, a loose structure with great porosity
220
is obtained (Figure S6c-f). This result can be explained that both the particle size of T-2 and T-1
221
are much smaller than that of T-3, thus, the amount of precipitated CaCO3 is relatively larger than
222
that required for the uniform coating of CaCO3 around the templates, as shown in Figure S7. This
223
is similar to the condition when sodium carbonate or oxalic acid is used as the precipitant, in which
224
the templates are embedded in precipitated CaCO3 or CaC2O4, subsequently leading to the
225
significant enhancement in porosity after the removal of templates via calcination.
226 227
Figure 5. SEM or TEM of CaO-based sorbents prepared from 0.2 M Ca(NO3)2 solution under
228
different conditions. (a): SEM image of CaO-U; (b): TEM image of CaO-U; (c): SEM image of
229
CaO-based sorbent prepared in the absence of urea; (d): SEM image of CaO-based sorbent
230
prepared without the presence of template (×1K); (e): High resolution SEM image of marked
231
area in (d) (×10K); (f): SEM image of CaO-based sorbent prepared using urea as the precipitant,
232
subsequently calcined at 800 °C directly for 1 h.
ACS Paragon Plus Environment
13
Environmental Science & Technology
Page 14 of 32
233
CO2 capture performance. Figure 6 and Table 2 show the cyclic carbonation conversions
234
of synthetic CaO-based sorbents as well as the reference limestone over the tested fifteen cycles
235
under both mild and harsh conditions, respectively. It is clear that all the synthetic CaO-based
236
sorbents show a significant enhancement in initial CO2 uptake compared to the reference
237
limestone, whether the hollow microsphere structure is formed. For example, the reference
238
limestone possesses an initial carbonation conversion of 74.5% and 63.2% under a mild and harsh
239
condition, respectively. While cage-like CaO hollow microspheres possess an initial one of 98.2%
240
and 82.5% under the corresponding conditions, exceeding that of the reference limestone by 31.8%
241
and 30.5%, respectively. The significant improvement in CO2 uptake capacity of the synthetic
242
CaO-based sorbents results from the higher specific surface area compared to the limestone (Table
243
S1), since surface area plays a critical role in CO2 capture performance of sorbents.12 Moreover, it
244
is well known that carbonation reaction consists of two stages: an initially rapid chemical-
245
controlled stage, followed by a substantially slower diffusion-controlled stage.7,8,14,49 As shown in
246
Figure 7a, a carbonation conversion of 69.4% can be obtain for cage-like CaO hollow
247
microspheres with only three min under a harsh condition, accounting for 84.1% of the total initial
248
CO2 uptake capacity. This result suggests cage-like hollow microsphere structure can significantly
249
improve the CO2 sorption kinetics, especially at the chemical-controlled stage.
250
During the tested fifteen repeated carbonation/calcination cycles, decays in CO2 uptake are
251
observed for all the sorbents. However, it should be noted that all the synthetic CaO-based sorbents
252
possess higher final CO2 uptake capacities compared to the limestone. Under a harsh condition,
253
cage-like CaO hollow microspheres possess a final carbonation conversion of 39.7%, exceeding
254
the reference limestone by 148.1%. While CaO-OA and CaO-SC have a final one of 22.8% and
255
18.8%, exceeding the reference limestone by 42.5% and 17.5%, respectively. And the
ACS Paragon Plus Environment
14
Page 15 of 32
Environmental Science & Technology
256
corresponding carbonation curves in Figure 7b confirm that all the synthetic CaO-based sorbents
257
still possess better CO2 sorption kinetics than the limestone. For the best material, i.e., cage-like
258
CaO hollow microspheres, an additional test of 30 repeated carbonation/calcination cycles under
259
a harsh condition is conducted. A carbonation conversion of 37.7% can be obtained after 30
260
repeated cycles (Figure S8), suggesting the good cyclic stability of CO2 capture performance for
261
cage-like CaO hollow microspheres. Besides, it is interesting to find that cage-like CaO hollow
262
microspheres possess better cyclic stability of CO2 capture than CaO-OA and CaO-SC. It is due
263
to not only the cage-like CaO hollow microspheres have unique hollow microsphere structure
264
compared to CaO-OA and CaO-SC, but also they can maintain this unique structure during the
265
repeated carbonation/calcination cycles even under a harsh condition. As clearly shown in Figure
266
8 and Figure S9, cage-like CaO hollow microspheres can maintain its unique hollow structure after
267
the repeated cycles, whereas the loose structures disappear and only compact surfaces with low
268
porosity are observed for both CaO-OA and CaO-SC.
269 270
Figure 6. CO2 capture performance of CaO-based sorbents and reference limestone over the
271
tested fifteen repeated cycles under (a) mild and (b) harsh conditions, respectively.
ACS Paragon Plus Environment
15
Environmental Science & Technology
Page 16 of 32
272
Table 2. Carbonation conversions of synthetic CaO-based sorbents and limestone under mild
273
and harsh condition Carbonation conversion sample
under a mild condition (%)
Carbonation conversion under a harsh condition (%)
1st cycle
15th cycle
1st cycle
15th cycle
Limestone
74.5
26.5
63.2
16.0
CaO-U
98.2
49.2
82.5
39.7
CaO-OA
85.1
33.9
71
22.8
CaO-SC
77.8
28.5
67.7
18.8
274
275 276
Figure 7. Carbonation curves of the synthetic CaO-based sorbents and the limestone under a
277
harsh condition in the (a): 1st cycle and (b): 15th cycle, respectively.
ACS Paragon Plus Environment
16
Page 17 of 32
Environmental Science & Technology
278 279
Figure 8. SEM images of (a): CaO-U, (b): CaO-OA, and (c): CaO-SC after the fifteen repeated
280
carbonation/calcination cycles under a mild condition.
281
Kinetic of carbonation reaction. In order to understand the reaction mechanism at different
282
stages in the carbonation process of cage-like CaO hollow microspheres, kinetic study using two
283
different models is conducted in this work. At the same time, the kinetic study of the reference
284
limestone is also done and used as the reference. Here the grain model and three-dimensional
285
diffusion model are used to build the relationship between carbonation conversion and time, which
286
are commonly used to analyze the reaction rate constants in chemical-controlled stage and
287
diffusion-controlled stage, respectively.50-54 The two models can be expressed as:
288
Grain model for chemical-controlled stage:
289
1- 1-Cn
13
kc t
(1)
290
Three-dimensional diffusion model for diffusion-controlled stage:
291 292
1- 1-Cn 1 3 kd t (2) where Cn is the carbonation conversion at the n cycle (in %), kc and kd are the reaction rate
293
constants during chemical-controlled stage and diffusion-controlled stage, respectively.
2
294
The results of kinetic calculation with models at two reaction stages for cage-like CaO hollow
295
microspheres and limestone are given in Figure 9. It can be found that the calculated data obtained
ACS Paragon Plus Environment
17
Environmental Science & Technology
Page 18 of 32
296
by kinetic models agree well with the fitting data, suggesting that both the grain model and the
297
three-dimensional diffusion model can be applied to analyze the two stages in the carbonation
298
process. The reaction rate constants of different stages can be obtained from the linear plot slope
299
and the results are given in Table S2. It can be found that the reaction rate constant of chemical-
300
controlled stage is always larger than that of diffusion-controlled stage, which agrees well with the
301
initial rapid increase of carbonation conversion followed by slow increase in Figure 7. And the
302
reaction rate constants of both chemical-controlled stage and diffusion-controlled stage in the 15th
303
cycle are lower than those in the 1st cycle, indicating the occurrence of decline in CO2 capture
304
performance. Moreover, it is interesting to find that the reaction rate constants for cage-like CaO
305
hollow microspheres are much higher than those of limestone, suggesting the good CO2 sorption
306
kinetics of cage-like CaO hollow microspheres. This can be related to the unique hollow
307
microsphere structure.
308 309
Figure 9. Kinetic calculation with two different models for cage-like CaO hollow
310
microspheres and limestone at chemical-controlled stage and diffusion-controlled stage in the 1st
311
and 15th cycles, respectively.
ACS Paragon Plus Environment
18
Page 19 of 32
312
Environmental Science & Technology
Here the kinetic parameters for carbonation conversion are calculated below by fitting the
313
reaction rate equation based on the grain model.50,53,54
314
The reaction rate of CaO-CO2 can be expressed as:
315 316
317 318
319 320
n dX 1 3 3r 1 X 56ks PCO 2 PCO 2,eq S dt 1 X At the initial time t = 0, Eq. (3) becomes:
R'
R'
n dX 3r0 56k s PCO 2 PCO 2,eq S0 dt
(3)
(4)
In logarithmic form, with ks A exp E R gT , Eq. (4) becomes:
56 E ln r0 n ln PCO 2 PCO 2,eq + ln S0 ln A 3 R gT
(5)
Assuming zero reaction order when PCO 2 PCO 2,eq 10kPa , Eq. (5) can be rearranged as:
322
56 E ln r0 l n AS0 (6) 3 R gT where r0 is the initial specific reaction rate in the carbonation reaction (in s-1), A is the pre-
323
exponential factor (in mol/(m2·s)), S0 is the initial surface area of sorbent (in m2/g), Rg is gas
324
constant (8.314 J/(mol·K)), T is carbonation temperature (in K), and E is the activation energy (in
325
kJ/mol). Thus, E and A can be determined by the slope and intercept of a plot of ln(r0) versus
326
1000/T, as shown in Figure 10. Considering the difficulty to determine the specific surface area of
327
sorbent before the 15th cycle because of the small amount of the sorbent sample for TGA test, the
328
kinetic parameters for both cage-like CaO hollow microspheres and limestone are calculated only
329
in the 1st cycle, as given in Table 3. The activation energy of limestone is determined as 39.2
330
kJ/mol, which is in the range obtained by other researchers in similar experimental conditions.50-
321
ACS Paragon Plus Environment
19
Environmental Science & Technology
Page 20 of 32
331
54
332
lower than that of limestone, which can be caused by the unique hollow microsphere structure.
It is interesting to find that the activation energy of cage-like CaO hollow microspheres is much
333 334
Figure 10. Arrhenius plots for carbonation at 550-750 °C in 15 vol.% CO2 (N2 balance) for (a):
335
cage-like CaO hollow microspheres and (b): limestone in the 1st cycle, respectively.
336
Table 3. Kinetic Parameters of the carbonation reaction of cage-like CaO hollow microspheres
337
and limestone sample
E (kJ/mol)
A (mol/(m2·s))
CaO-U
28.1
4.7×10-3
Limestone
39.2
0.12
338 339
Life cycle assessment. A “cradle-to-factory gate” LCA approach is used in this work, which
340
has been reported in previous work.55,56 Figure 11 outlines the system boundary for this LCA study.
341
It should be noted that some parameters are not considered due to the lack of source data of relevant
ACS Paragon Plus Environment
20
Page 21 of 32
Environmental Science & Technology
342
chemicals/processes. Therefore, it should be treated with caution when applying for other synthetic
343
processes or materials.
344 345
Figure 11. System boundary for LCA of cage-like CaO hollow microspheres via a template-
346
assisted synthesis approach. There are three main steps (1-3) in the material synthesis.
347
The material and energy consumption data during the synthesis are summarized in Tables S3
348
-S5, respectively. Therefore, we can obtain the production cost distribution of raw materials,
349
electricity and biomass (as fuel) in each step for the synthesis of one kg cage-like CaO hollow
350
microspheres according to Tables S3-S5, as shown in Figure 12. The calculated production cost
351
for synthesis of one kg cage-like CaO hollow microspheres is 24.95 RMB, 86.1% of which is spent
352
on raw material. Although this production cost seems high, it should be noted that the cost here is
353
based on our laboratory data, and can be reduced by a large extent by scaling up. For example,
354
recycling the chemicals such as urea and Ca(NO3)2 is important to reduce the production cost,
355
since these two raw materials account for 51.4% of the total cost. According to the material
356
balance, 80% of the urea and 10.7% of the Ca2+ ions can be recycled. Moreover, Ca(NO3)2 can be
357
replaced by other cheap calcium based material like carbide slag.51 The use of CS templates
ACS Paragon Plus Environment
21
Environmental Science & Technology
Page 22 of 32
358
derived from the hydrothermal treatment of other cheap starting material such as glucose41 will
359
also largely reduce the production cost. And the heat derived from the removal of CS templates
360
via combustion process can be recovered in a large-scale synthesis, which can further lower the
361
production cost. Through all the improvement, the cost of one kg cage-like CaO hollow
362
microspheres can be reduced to 7.53 RMB.
363 364
Figure 12. Production cost distribution of raw materials, electricity and biomass in each step for
365
the synthesis of one kg cage-like CaO hollow microspheres.
366
The unit price of limestone is 700 RMB/t, which is lower than production cost of CaO hollow
367
microspheres. However, the significant decline in CO2 capture performance of limestone should
368
be taken into account when conducting the economic analysis. In terms of limestone, a relative
369
makeup (fresh limestone/sorbent circulation) rate is generally determined as 0.04, as reported in
370
previous work.57,58 As a result, for one kg limestone, the overall CO2 captured can be calculated
371
as 2.6 kg based on its capacity curve of the raw limestone over recycle numbers. Whereas this
372
makeup rate can be greatly reduced when using our sorbent, due to the significant enhanced CO2
ACS Paragon Plus Environment
22
Page 23 of 32
Environmental Science & Technology
373
capture performance of cage-like CaO hollow microspheres. The carbonation conversion rate of
374
our sorbent can be stable at 37.7% after 30 cycles under a harsh condition, while the carbonation
375
conversion rate of the parent limestone is only 16% after 15 cycles. In this case, the decreased cost
376
of CO2 capture for cage-like CaO hollow microspheres can offset the production cost by reducing
377
the relative makeup rate. The overall cost of cage-like CaO hollow microspheres is expected to be
378
lower than the raw limestone capturing the same amount of CO2. Furthermore, biomass is used to
379
supply the heat that required for the synthesis process. And the templates used in this work are
380
derived from biomass. Therefore, the synthesis of cage-like CaO hollow microspheres is carbon
381
neutral in nature.
382
ASSOCIATED CONTENT
383
Supporting Information.
384
This information is available free of charge via the Internet at http://pubs.acs.org.
385
Experimental preparation of CS templates; Experimental synthesis of CaO-based sorbents;
386
Experimental characterization; Experimental performance test; SEM images of the CS templates
387
prepared under different hydrothermal duration; SEM images of the CS templates prepared under
388
different hydrothermal temperature; Raman spectra of T-3; SEM images of T-3 before and after
389
Ca2+ ions deposition in the presence of urea; High-resolution SEM image of cage-like CaO hollow
390
microspheres prepared using T-3 as template and from 0.5 M Ca(NO3)2 solution; SEM images of
391
CaO-based sorbents using T-3, T-2 and T-1 as the template; SEM images of CaO-based sorbents
392
using T-2 and T-1 as the template before template removal process; Cyclic carbonation conversion
393
of cage-like CaO hollow microspheres over the tested 30 cycles under a harsh condition; TEM
394
image of cage-like CaO hollow microspheres after fifteen cycles under a harsh condition; TEM
ACS Paragon Plus Environment
23
Environmental Science & Technology
Page 24 of 32
395
image of cage-like CaO hollow microspheres after fifteen repeated carbonation/calcination cycles
396
under a harsh condition; Reaction rate constant during chemical-controlled and diffusion-
397
controlled stages in the 1st cycle and 15th cycle carbonation process of CaO hollow microspheres
398
and limestone; material, electricity and biomass consumptions for the synthesis of cage-like CaO
399
hollow microspheres.
400 401
AUTHOR INFORMATION
402
Corresponding Author
403
*Email:
[email protected] 404
Author Contributions
405
The manuscript was written through contributions of all authors. All authors have given
406
approval to the final version of the manuscript.
407
Funding Sources
408
The authors wish to acknowledge the financial support from the National Key Research and
409
Development Program of China (2016YFE0102500-06-01).
410
Notes
411 412 413
The authors declare no competing financial interest. ACKNOWLEDGMENT The authors wish to acknowledge the financial support from the National Key Research and
414
Development Program of China (2016YFE0102500-06-01).
415
REFERENCES
ACS Paragon Plus Environment
24
Page 25 of 32
Environmental Science & Technology
416
(1) Antzara, A.; Heracleous, E.; Lemonidou, A. A. Improving the Stability of Synthetic CaO-
417
Based CO2 Sorbents by Structural Promoters. Appl. Energy 2015, 156, 331–343.
418
(2) Xu, Y.; Ding, H.; Luo, C.; Zheng, Y.; Zhang, Q.; Li, X.; Sun, J.; Zhang, L. Potential Synergy
419
of Chlorine and Potassium and Sodium Elements in Carbonation Enhancement of CaO-Based
420
Sorbents. ACS Sustain. Chem. Eng. 2018, 6 (9), 11677–11684.
421
(3) Boot-Handford, M. E.; Abanades, J. C.; Anthony, E. J.; Blunt, M. J.; Brandani, S.; Mac
422
Dowell, N.; Fernández, J. R.; Ferrari, M. C.; Gross, R.; Hallett, J. P.; et al. Carbon Capture and
423
Storage Update. Energy Environ. Sci. 2014, 7 (1), 130–189.
424
(4) Xu, Y.; Ding, H.; Luo, C.; Zheng, Y.; Xu, Y.; Li, X.; Zhang, Z.; Shen, C.; Zhang, L. Effect of
425
Lignin, Cellulose and Hemicellulose on Calcium Looping Behavior of CaO-Based Sorbents
426
Derived from Extrusion-Spherization Method. Chem. Eng. J. 2018, 334 (11), 2520–2529.
427
(5) Bhown, A. S.; Freeman, B. C. Analysis and Status of Post-Combustion Carbon Dioxide
428
Capture Technologies. Environ. Sci. Technol. 2011, 45 (20), 8624–8632.
429
(6) Kierzkowska, A. M.; Pacciani, R.; Müller, C. R. CaO-Based CO2 Sorbents: From
430
Fundamentals to the Development of New, Highly Effective Materials. ChemSusChem 2013, 6
431
(7), 1130–1148.
432
(7) Yan, F.; Jiang, J.; Li, K.; Liu, N.; Chen, X.; Gao, Y.; Tian, S. Green Synthesis of Nanosilica
433
from Coal Fly Ash and Its Stabilizing Effect on CaO Sorbents for CO2 Capture. Environ. Sci.
434
Technol. 2017, 51 (13), 7606–7615.
ACS Paragon Plus Environment
25
Environmental Science & Technology
Page 26 of 32
435
(8) Peng, W.; Xu, Z.; Luo, C.; Zhao, H. Tailor-Made Core-Shell CaO/TiO2-Al2O3 Architecture
436
as a High-Capacity and Long-Life CO2 Sorbent. Environ. Sci. Technol. 2015, 49 (13), 8237–
437
8245.
438
(9) Tian, S.; Jiang, J.; Yan, F.; Li, K.; Chen, X. Synthesis of Highly Efficient CaO-Based, Self-
439
Stabilizing CO2 Sorbents via Structure-Reforming of Steel Slag. Environ. Sci. Technol. 2015, 49
440
(12), 7464–7472.
441
(10) Criado, Y. A.; Arias, B.; Abanades, J. C. Calcium Looping CO2 capture System for Back-up
442
Power Plants. Energy Environ. Sci. 2017, 10 (9), 1994–2004.
443
(11) Abanades, J. C.; Grasa, G.; Alonso, M.; Rodriguez, N.; Anthony, E. J.; Romeo, L. M. Cost
444
Structure of a Postcombustion CO2 Capture System Using CaO. Environ. Sci. Technol. 2007, 41
445
(15), 5523–5527.
446
(12) Chen, J.; Duan, L.; Donat, F.; Müller, C. R.; Anthony, E. J.; Fan, M. Self-Activated ,
447
Nanostructured Composite for Improved CaL-CLC Technology. Chem. Eng. J. 2018, 351 (6),
448
1038–1046.
449
(13) Broda, M.; Manovic, V.; Anthony, E. J.; Müller, C. R. Effect of Pelletization and Addition
450
of Steam on the Cyclic Performance of Carbon-Templated, CaO-Based CO2 Sorbents. Environ.
451
Sci. Technol. 2014, 48 (9), 5322–5328.
452
(14) Wang, S.; Fan, S.; Fan, L.; Zhao, Y.; Ma, X. Effect of Cerium Oxide Doping on the
453
Performance of CaO-Based Sorbents during Calcium Looping Cycles. Environ. Sci. Technol.
454
2015, 49 (8), 5021–5027.
ACS Paragon Plus Environment
26
Page 27 of 32
Environmental Science & Technology
455
(15) Donat, F.; Florin, N. H.; Anthony, E. J.; Fennell, P. S. Influence of High-Temperature
456
Steam on the Reactivity of CaO Sorbent for CO2 Capture. Environ. Sci. Technol. 2012, 46 (2),
457
1262–1269.
458
(16) Manovic, V.; Anthony, E. J. Thermal Activation of CaO-Based Sorbent and Self-
459
Reactivation during CO2 Capture Looping Cycles. Environ. Sci. Technol. 2008, 42 (11), 4170–
460
4174.
461
(17) Sun, P.; Grace, J. R.; Lim, C. J.; Anthony, E. J. Investigation of Attempts to Improve Cyclic
462
CO2 capture by Sorbent Hydration and Modification. Ind. Eng. Chem. Res. 2008, 47 (6), 2024–
463
2032.
464
(18) Manovic, V.; Anthony, E. J.; Grasa, G.; Abanades, J. C. CO2 looping Cycle Performance of
465
a High-Purity Limestone after Thermal Activation/Doping. Energy and Fuels 2008, 22 (5),
466
3258–3264.
467
(19) Broda, M.; Kierzkowska, A. M.; Müller, C. R. Influence of the Calcination and Carbonation
468
Conditions on the CO2 Uptake of Synthetic Ca-Based CO2 Sorbents. Environ. Sci. Technol.
469
2012, 46 (19), 10849–10856.
470
(20) Kierzkowska, A. M.; Pacciani, R.; Müller, C. R. CaO-Based CO2 Sorbents: From
471
Fundamentals to the Development of New, Highly Effective Materials. ChemSusChem 2013, 6
472
(7), 1130–1148.
473
(21) Erans, M.; Manovic, V.; Anthony, E. J. Calcium Looping Sorbents for CO2 capture. Appl.
474
Energy 2016, 180, 722–742.
ACS Paragon Plus Environment
27
Environmental Science & Technology
Page 28 of 32
475
(22) Liu, W.; Low, N. W. L.; Feng, B.; Wang, G.; Diniz da Costa, J. C. Calcium Precursors for
476
the Production of CaO Sorbents for Multicycle CO2 Capture. Environ. Sci. Technol. 2010, 44 (2),
477
841–847.
478
(23) Naeem, M. A.; Armutlulu, A.; Imtiaz, Q.; Donat, F.; Schäublin, R.; Kierzkowska, A.;
479
Müller, C. R. Optimization of the Structural Characteristics of CaO and Its Effective
480
Stabilization Yield High-Capacity CO2 sorbents. Nat. Commun. 2018, 9 (1), 1–11.
481
(24) Li, Y.; Shi, J. Hollow-Structured Mesoporous Materials: Chemical Synthesis,
482
Functionalization and Applications. Adv. Mater. 2014, 26 (20), 3176–3205.
483
(25) Zhou, L.; Zhuang, Z.; Zhao, H.; Lin, M.; Zhao, D.; Mai, L. Intricate Hollow Structures:
484
Controlled Synthesis and Applications in Energy Storage and Conversion. Adv. Mater. 2017, 29
485
(20).
486
(26) Yu, L.; Hu, H.; Wu, H. Bin; Lou, X. W. D. Complex Hollow Nanostructures: Synthesis and
487
Energy-Related Applications. Adv. Mater. 2017, 29 (15).
488
(27) Ping, H.; Wu, S. Preparation of Cage-like Nano-CaCO3 Hollow Spheres for Enhanced CO2
489
Sorption. RSC Adv. 2015, 5 (80), 65052–65057.
490
(28) Derevschikov, V.; Semeykina, V.; Bitar, J.; Parkhomchuk, E.; Okunev, A. Template
491
Technique for Synthesis of CaO-Based Sorbents with Designed Macroporous Structure.
492
Microporous Mesoporous Mater. 2017, 238, 56–61.
493
(29) Liu, F.-Q.; Li, W.-H.; Liu, B.-C.; Li, R.-X. Synthesis, Characterization, and High
494
Temperature CO2 Capture of New CaO Based Hollow Sphere Sorbents. J. Mater. Chem. A 2013,
495
1 (27), 8037.
ACS Paragon Plus Environment
28
Page 29 of 32
Environmental Science & Technology
496
(30) Naeem, M. A.; Armutlulu, A.; Broda, M.; Lebedev, D.; Müller, C. R. The Development of
497
Effective CaO-Based CO2 Sorbents via a Sacrificial Templating Technique. Faraday Discuss.
498
2016, 192, 85–95.
499
(31) Yu, X. F.; Liu, Y. H.; Zhou, Z. W.; Xiong, G. X.; Cao, X. H.; Li, M.; Zhang, Z. Bin.
500
Adsorptive Removal of U(VI) from Aqueous Solution by Hydrothermal Carbon Spheres with
501
Phosphate Group. J. Radioanal. Nucl. Chem. 2014, 300 (3), 1235–1244.
502
(32) Sun, X.; Li, Y. Colloidal Carbon Spheres and Their Core/Shell Structures with Noble-Metal
503
Nanoparticles. Angew. Chemie - Int. Ed. 2004, 43 (5), 597–601.
504
(33) Yu, J.; Yu, X.; Huang, B.; Zhang, X.; Dai, Y. Hydrothermal Synthesis and Visible-Light
505
Photocatalytic Activity of Novel Cage-like Ferric Oxide Hollow Spheres. Cryst. Growth Des.
506
2009, 9 (3), 1474–1480.
507
(34) Wang, Q.; Li, H.; Chen, L.; Huang, X. Monodispersed Hard Carbon Spherules with
508
Uniform Nanopores. Carbon N. Y. 2001, 39 (14), 2211–2214.
509
(35) Sevilla, M.; Fuertes, A. B. Chemical and Structural Properties of Carbonaceous Products
510
Obtained by Hydrothermal Carbonization of Saccharides. Chem. - A Eur. J. 2009, 15 (16), 4195–
511
4203.
512
(36) Zhu, J.; Li, R.; Niu, W.; Wu, Y.; Gou, X. Facile Hydrogen Generation Using Colloidal
513
Carbon Supported Cobalt to Catalyze Hydrolysis of Sodium Borohydride. J. Power Sources
514
2012, 211, 33–39.
515
(37) Deng, L.; Young, R. J.; Kinloch, I. A.; Zhu, Y.; Eichhorn, S. J. Carbon Nanofibres Produced
516
from Electrospun Cellulose Nanofibres. Carbon N. Y. 2013, 58 (0), 66–75.
ACS Paragon Plus Environment
29
Environmental Science & Technology
Page 30 of 32
517
(38) Bystrzejewski, M.; Karoly, Z.; Szepvolgyi, J.; Kaszuwara, W.; Huczko, A.; Lange, H.
518
Continuous Synthesis of Carbon-Encapsulated Magnetic Nanoparticles with a Minimum
519
Production of Amorphous Carbon. Carbon N. Y. 2009, 47 (8), 2040–2048.
520
(39) Armutlulu, A.; Naeem, M. A.; Liu, H.-J.; Kim, S. M.; Kierzkowska, A.; Fedorov, A.;
521
Müller, C. R. Multishelled CaO Microspheres Stabilized by Atomic Layer Deposition of Al2O3
522
for Enhanced CO2 Capture Performance. Adv. Mater. 2017, 29 (41), 1702896.
523
(40) Zhenmin, L.; Xiaoyong, L.; Hong, W.; Dan, M.; Chaojian, X.; Dan, W. General Synthesis
524
of Homogeneous Hollow Core-Shell Ferrite Microspheres. J. Phys. Chem. C 2009, 113 (7),
525
2792–2797.
526
(41) Sun, X.; Liu, J.; Li, Y. Use of Carbonaceous Polysaccharide Microspheres as Templates for
527
Fabricating Metal Oxide Hollow Spheres. Chem. - A Eur. J. 2006, 12 (7), 2039–2047.
528
(42) Sevilla, M.; Fuertes, A. B. Chemical and Structural Properties of Carbonaceous Products
529
Obtained by Hydrothermal Carbonization of Saccharides. Chem. - A Eur. J. 2009, 15 (16), 4195–
530
4203.
531
(43) Ferrari, A. C.; Robertson, J. Interpretation of Raman Spectra of Disordered and Amorphous
532
Carbon. Phys. Rev. B 2000, 61 (20), 14095–14107.
533
(44) Schwan, J.; Ulrich, S.; Batori, V.; Ehrhardt, H.; Silva, S. R. P. Raman Spectroscopy on
534
Amorphous Carbon Films. J. Appl. Phys. 1996, 80 (1), 440–447.
535
(45) Duan, L.; Su, C.; Erans, M.; Li, Y.; Anthony, E. J.; Chen, H. CO2 Capture Performance
536
Using Biomass-Templated Cement-Supported Limestone Pellets. Ind. Eng. Chem. Res. 2016, 55
537
(39), 10294–10300.
ACS Paragon Plus Environment
30
Page 31 of 32
Environmental Science & Technology
538
(46) Ridha, F. N.; Wu, Y.; Manovic, V.; Macchi, A.; Anthony, E. J. Enhanced CO2 capture by
539
Biomass-Templated Ca(OH)2-Based Pellets. Chem. Eng. J. 2015, 274, 69–75.
540
(47) Koebel, M.; Strutz, E. O. Thermal and Hydrolytic Decomposition of Urea for Automotive
541
Selective Catalytic Reduction Systems: Thermochemical and Practical Aspects. Ind. Eng. Chem.
542
Res. 2003, 42 (10), 2093–2100.
543
(48) Sun, X.; Li, Y. Ga2O3 and GaN Semiconductor Hollow Spheres. Angew. Chemie - Int. Ed.
544
2004, 43 (29), 3827–3831.
545
(49) Valverde, J. M.; Perejon, A.; Perez-Maqueda, L. A. Enhancement of Fast CO2 capture by a
546
Nano-SiO2/CaO Composite at Ca-Looping Conditions. Environ. Sci. Technol. 2012, 46 (11),
547
6401–6408.
548
(50) Zhang, Y.; Gong, X.; Chen, X.; Yin, L.; Zhang, J.; Liu, W. Performance of Synthetic CaO-
549
Based Sorbent Pellets for CO2 capture and Kinetic Analysis. Fuel 2018, 232 (7), 205–214.
550
(51) Li, Y.; Liu, H.; Sun, R.; Wu, S.; Lu, C. Thermal Analysis of Cyclic Carbonation Behavior
551
of CaO Derived from Carbide Slag at High Temperature. J. Therm. Anal. Calorim. 2012, 110
552
(2), 685–694.
553
(52) Sun, P.; Grace, J. R.; Lim, C. J.; Anthony, E. J. Determination of Intrinsic Rate Constants of
554
the CaO-CO2 reaction. Chem. Eng. Sci. 2008, 63 (1), 47–56.
555
(53) Symonds, R. T.; Lu, D. Y.; Hughes, R. W.; Anthony, E. J.; Macchi, A. CO2 capture from
556
Simulated Syngas via Cyclic Carbonation/Calcination for a Naturally Occurring Limestone:
557
Pilot-Plant Testing. Ind. Eng. Chem. Res. 2009, 48 (18), 8431–8440.
ACS Paragon Plus Environment
31
Environmental Science & Technology
Page 32 of 32
558
(54) Yin, J.; Qin, C.; Feng, B.; Ge, L.; Luo, C.; Liu, W.; An, H. Calcium Looping for CO2 Capture
559
at a Constant High Temperature. Energy and Fuels 2014, 28 (1), 307–318.
560
(55) Gao, T.; Jelle, B. P.; Sandberg, L. I. C.; Gustavsen, A. Monodisperse Hollow Silica
561
Nanospheres for Nano Insulation Materials: Synthesis, Characterization, and Life Cycle
562
Assessment. ACS Appl. Mater. Interfaces 2013, 5 (3), 761–767.
563
(56) Deorsola, F. A.; Russo, N.; Blengini, G. A.; Fino, D. Synthesis, Characterization and
564
Environmental Assessment of Nanosized MoS2 particles for Lubricants Applications. Chem. Eng.
565
J. 2012, 195–196, 1–6.
566
(57) Hanak, D. P.; Anthony, E. J.; Manovic, V. A Review of Developments in Pilot-Plant Testing
567
and Modelling of Calcium Looping Process for CO2 Capture from Power Generation Systems.
568
Energy Environ. Sci. 2015, 8 (8), 2199–2249.
569
(58) Hanak, D. P.; Erans, M.; Nabavi, S. A.; Jeremias, M.; Romeo, L. M.; Manovic, V. Technical
570
and Economic Feasibility Evaluation of Calcium Looping with No CO2 recirculation. Chem. Eng.
571
J. 2018, 335, 763–773.
ACS Paragon Plus Environment
32