Subscriber access provided by UNIVERSITY OF ADELAIDE LIBRARIES
Article
Demethanation Trend of Hydrochar Induced by Organic Solvent Washing and its Influence on Hydrochar Activation Xiangdong Zhu, Yuchen Liu, Feng Qian, Zhongfang Lei, Zhenya Zhang, Shicheng Zhang, Jianmin Chen, and Zhiyong Jason Ren Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.6b06594 • Publication Date (Web): 20 Aug 2017 Downloaded from http://pubs.acs.org on August 21, 2017
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 free 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 accessible to all readers and 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.
Environmental Science & Technology 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 29
Environmental Science & Technology
ACS Paragon Plus Environment
Environmental Science & Technology
1
Demethanation Trend of Hydrochar Induced by Organic Solvent
2
Washing and its Influence on Hydrochar Activation
3
Xiangdong Zhu,†,‡ Yuchen Liu,† Feng Qian,† Zhongfang Lei,# Zhenya Zhang,# Shicheng
4
Zhang,†,* Jianmin Chen,† Zhiyong Jason Ren‡
5
†
6
Environmental Science and Engineering, Fudan University, Shanghai 200433, China
7
‡
8
Boulder, CO 80309, United States
Shanghai Key Laboratory of Atmospheric Particle Pollution and Prevention (LAP3), Department of
Department of Civil, Environmental, and Architectural Engineering, University of Colorado Boulder,
9 10
# Graduate School of Life and Environmental Sciences, University of Tsukuba, Tsukuba 305-8572, Japan
11 12
* Corresponding author, Tel/fax: +86-21-65642297; E-mail:
[email protected] (Shicheng
13
Zhang).
14
1 ACS Paragon Plus Environment
Page 2 of 29
Page 3 of 29
15
Environmental Science & Technology
ABSTRACT
16
Hydrochar derived from hydrothermal carbonization (HTC) has been recognized as a
17
promising carbonaceous material for environmental remediation. Organic solvents are widely
18
used to extract bio-oil from hydrochar after HTC, but the effects of solvent extraction on
19
hydrochar characteristics have not been investigated. This study comprehensively analyzed the
20
effects of different washing times and solvent types on hydrochar properties. Results indicate that
21
the weight loss of hydrochar by tetrahydrofuran washing occurred mainly in the first 90 min, and
22
the loss ratios of elements followed a descending order of H > C > O, resulting in a decrease in
23
H/C atomic ratio while an increase in O/C atomic ratio. The use of various solvents for washing
24
brought about hydrochar loss ratios in a descending order of petroleum ether < dichloromethane
25
< acetone < tetrahydrofuran. Results from the Van Krevelen diagram, FTIR, 13C NMR and XPS
26
further confirmed that demethanation controlled this washing process. Most importantly, the
27
surface area of hydrochar increased after bio-oil removal via washing, which promoted the
28
surface area development for hydrochar-derived magnetic carbon composites, but to some extent
29
decreased the microporosity. Additionally, hydrochar washing by organic solvent has important
30
implications for the global carbon cycle and its remediation application.
31
2 ACS Paragon Plus Environment
Environmental Science & Technology
32 33
Page 4 of 29
INTRODUCTION Compared with the slow pyrolysis for the production of biochar,1,
2
hydrothermal
34
carbonization (HTC) is an emerging process for simultaneous production of renewable bio-oil
35
and hydrochar material from biomass.3 The HTC conversion of biomass demonstrated great
36
potentials for renewable energy generation and environmental remediation.4-6 Hydrochar
37
generated from HTC is an excellent carbonaceous material due to its distinct properties including
38
high carbon (C) content, thermal stability, and moderate porosity.7, 8 Extensive studies have been
39
performed to regulate and optimize the properties of hydrochar so its performance can be
40
improved for targeted applications especially in the field of environmental remediation and
41
agricultural land application.9-12
42
It is well characterized that the surface of hydrochar shows strong hydrophobicity due to the
43
presence of bio-oil. Therefore, organic solvents such as acetone (AT) and tetrahydrofuran (THF)
44
are often used to wash hydrochar for bio-oil (also known as heavy oil) collection.13, 14 Previous
45
studies reported that bio-oil has low oxygen (O) content, while
46
O-containing groups of char were enhanced after methanol washing.17 These findings imply that
47
the loss ratios of C, O and hydrogen (H) contents in hydrochar may be different during washing,
48
because if C, H and O have the same loss ratios, O-containing groups of char would not change.
49
Therefore, it is reasonable to hypothesize that the properties of hydrochar like ratios of H/C
50
(relating to aromaticity) and O/C (relating to polarity) are would be altered after solvent washing,
51
which can greatly affect the char’s property for following applications.9, 17 However, there is little 3 ACS Paragon Plus Environment
15, 16
others reported that the
Page 5 of 29
Environmental Science & Technology
52
information on how solvent-based washing influences hydrochar properties. For example,
53
hydrochar is often used as a raw material in the preparation of high-porosity magnetic carbon
54
composites (MCs).18 It has been well documented that the thermal stability of hydrochar
55
(indicated by its H/C atomic ratio) strongly affects the porosity of hydrochar-derived MCs.9, 19, 20
56
Solvent washing can change the H/C atomic ratio as the loss ratios of C and H contents can be
57
different, so the porosity of the hydrochar-derived MCs can vary. On the other hand, the porosity
58
of hydrochar will also change after the loss of bio-oil from the hydrochar surface. As a result,
59
more chemical activators can be loaded into the pore system, which helps to activate the
60
hydrochar material with higher porosity. However, the connection between the properties of
61
solvent-washed hydrochar and the porosity of hydrochar-derived MCs has not yet been
62
established.
63
To investigate this unknown mechanism, the aim of this study is to analyze the effects of
64
washing times and solvent types on the changes of hydrochar characteristics, such as elemental
65
composition, loss ratios of elemental contents (C, H, O and nitrogen, N), and changes in porosity.
66
Specifically, several tasks were performed including (i) the role of washing time on the
67
hydrochar washing using THF solvent, (ii) the evaluation of solvent types on hydrochar washing,
68
(iii) data analysis and interpretation of the characteristics of hydrochar washing using the Van
69
Krevelen diagram, Fourier transform infrared spectroscopy (FTIR), carbon nuclear magnetic
70
resonance (13C NMR), and X-ray photoelectron spectroscopy (XPS) techniques, and (iv) drawing
71
the connection between the properties of washed hydrochar and the properties of
4 ACS Paragon Plus Environment
Environmental Science & Technology
72
hydrochar-derived MCs in terms of porosity, magnetic composition, and acid resistance.
73
EXPERIMENTAL SECTION
74
Chemicals and Materials
75
Hydrochar was obtained from the solid residual of Salix psammophila (SP) HTC reaction.
76
Typically, 44 L of water and 3.5 kg of SP were reacted in an autoclave, which was heated up to
77
280 °C and kept with 60 min. When the reactor was cooled to room temperature by tap water, the
78
hydrochar samples were collected by filtration.21 Four analytical grade organic solvents,
79
including petroleum ether (PE), dichloromethane (DLM), acetone (AT), and THF, were applied
80
for hydrochar washing for bio-oil removal. Zinc chloride (ZnCl2) and iron(III) chloride (FeCl3)
81
were applied to modify the washed hydrochar. Reagent grade bisphenol A (BPA) was purchased
82
from Aladdin Reagent Corporation.
83
Hydrochar Washing Procedure
84
An ultrasonic treatment was used in the hydrochar washing with a duration of 30 min each.
85
The ratio of hydrochar weight (g) to solvent volume (mL) was kept at 1:10. After washing, the
86
hydrochar was filtered and re-washed under the same condition. Considering the practical
87
extraction of bio-oil from hydrochar, the hydrochar was washed with 5 times using THF for a
88
total duration of 150 min. The washed hydrochar is denoted as H-X, where X represents the
89
washing time. To examine the effects of different solvent types, PE, DLM, AE, and THF were
90
applied with washing for 3 times each. The hydrochar was also washed with water (WT) as a
91
control experiment. The washed hydrochar is denoted as H-Y, where Y represents the solvent 5 ACS Paragon Plus Environment
Page 6 of 29
Page 7 of 29
Environmental Science & Technology
92
used for hydrochar washing. Before characterization measurements, the washed hydrochar was
93
heated at 80 °C to constant weight for the removal of residual organic solvent.
94
Modification and Characterization of Samples
95
Washed hydrochar was modified for MC synthesis using a simultaneous activation and
96
magnetization method previously reported.22 Briefly, 6 g of ZnCl2, 1.95 g of FeCl3 and 6 g of
97
washed hydrochar were mixed in 30 mL of water. The mixture was then shaken for 24 h and
98
air-dried at 80 °C for 4 h. Subsequently, the dried mixture was heat-treated at 600 °C for 90 min
99
under a nitrogen gas (N2) flow of 1 L/min. The resultant sample was successively washed with
100
0.1 M HCl and water and dried at 100 °C for 4 h. The prepared MCs are henceforth denoted as
101
MC-X, where X represents the solvent used for hydrochar washing.
102
The elemental composition of samples was analyzed with an elemental analyzer (Vario EL III,
103
Germany). The porosity of samples was determined by N2 adsorption at 77 K using a Quantasorb
104
instrument. The BET method was used to calculate the surface area (SBET) based on the partial
105
pressure (P/P0) range from 0.04 to 0.2, the total pore volume (Vt) was obtained from the
106
adsorbed nitrogen amount at a relative pressure of 0.99. The micropore volume (Vmic) and
107
micropore surface area (Smic) was determined by t-plot analysis. CO2 adsorption isotherms were
108
performed by using a Quantachrome FL33426 at 273 K and used to determine micropore volume,
109
micropore surface area and narrow micropore size distribution (< 1.5 nm) through the fitting of
110
DFT model.
111
The functional groups were examined using FTIR,
13
C CP/MAS NMR and XPS. Fourier
6 ACS Paragon Plus Environment
Environmental Science & Technology
Page 8 of 29
112
transform infrared spectroscopy (FTIR, Nexus470) was performed with a wavenumber range of
113
4000-400 cm-1 at a resolution of 2 cm-1. X-ray photoelectron spectroscopy (XPS) was examined
114
on a RBD-upgraded PHI-5000C ESCA system (Perkin Elmer). The X-ray anode was carried at
115
250 W and the voltage was maintained at 14.0 kV with a detection angle at 54°. Binding energies
116
were calibrated by setting C 1s at 284.6 eV. Solid state 13C nuclear magnetic resonance (NMR)
117
spectra with cross polarization magic angle spinning (CPMAS) were acquired on a Bruker DSX
118
300 NMR spectrometer with 7 mm zirconia rotors. The spinning speed of 4.8 kHz was used, and
119
8192 data points were collected. 1H t1 relaxation time and contact time was set to 2 s and 2.5 ms.
120
The phase structure of samples was characterized by X-ray diffraction (XRD) using Cu K α
121
radiation (λ = 1.5406 Å) at a scan rate of 8°/min and a step size of 0.02° in 2θ. The morphology
122
of samples was examined through a scanning electron microscopy (SEM, TS 5136MM). A
123
thermogravimetry (TG) analyzer was used to examine the stability of hydrochar, ~ 20 mg
124
samples that were heated from 30 to 800 °C in an N2 atmosphere at a rate of 20 °C/min.
125
Mössbauer spectroscopy was performed at room temperature with a cobalt isotope
126
source using 25 µm α-Fe foil as a reference. The Mössbauer spectroscopy of samples was fitted
127
with two components (Fe3O4 and ZnFe2O4) by the standard least square method. The acid
128
resistance of MCs (to test Fe leaching) was achieved at an MC concentration of 2000 mg/L under
129
a pH of 3.0 with a 24 h contact time. The extracted Fe solution was determined by inductively
130
coupled plasma (ICP, P-4010).
131
The
composition
of
extracted
bio-oil
was
qualitatively
7 ACS Paragon Plus Environment
determined
57
Co(Pd)
using
gas
Page 9 of 29
Environmental Science & Technology
132
chromatography/mass spectrometry (GC-MS) (Thermo FOCUS DSQ) with a HP-5 ms column.
133
The GC column was programmed to heat at a rate of 30 °C/min from 60 °C (held for 2 min) to
134
300°C (held for 5 min).
135
The adsorption isotherms of BPA onto MCs were achieved in the range of 10 - 200 mg/L BPA
136
(20% methanol) at 25 °C to test the adsorption ability of MCs. The concentration of the MC
137
sample was 200 mg/L. The concentrations of BPA were determined by high-performance liquid
138
chromatography (HPLC) at 280 nm with methanol and ultrapure water as a mobile phase at a
139
volume ratio of 75:25.
140
RESULTS AND DISCUSSION
141
Dynamic Changes in Hydrochar Characteristics Under Different Washing Times
142
As shown in Table 1, the loss of hydrochar mass and C, H, O and N contents mainly occurred
143
in the first three THF extractions, while only slight changes were observed thereafter.
144
Intraparticle diffuse model confirmed that the bio-oil adsorbed onto the exterior surface of
145
hydrochar was first washed off, which was followed by the extraction of bio-oil adsorbed onto
146
the interior hydrochar surface (Figure S1a). It is interesting to note that the C and H contents of
147
hydrochar decreased, but the O and N contents of hydrochar increased after THF extraction,
148
indicating that the loss ratios of C and H contents were notably higher than those of O and N
149
contents. As shown in Figure S1b, the loss behaviors of hydrochar due to THF washing can be
150
described well by the pseudo-second-order model, suggesting that binuclear desorption
151
mechanism regulated on the hydrochar washing by THF. Obviously, the loss of N content had the 8 ACS Paragon Plus Environment
Environmental Science & Technology
Page 10 of 29
152
fastest rate in the hydrochar washing by THF, indicating the N element of hydrochar appeared
153
most sensitive to THF extracting. The equation of pseudo-second-order model could be observed
154
at note of Figure S1b.
155
Small molecules including aromatic phenol, aliphatic acid and aliphatic alcohol were detected
156
in the GC-MS chromatograph of the THF-extracted bio-oil (Figure S2 and Table S1).
157
Interestingly, these compounds haven’t been detected after the second THF washing, indicating
158
that they were completely extracted prior to the first two washings. However, weight loss of
159
hydrochar continued. These results suggest that THF-soluble polymers like fragments of biomass
160
components and restructured organic matters were undetectable in the GC-MS analysis and
161
possessed a lower loss ratio resulted from THF washing.
162
Due to the pore blockage of hydrophobic compounds, the hydrochar possessed an extremely
163
low surface area (Table S2). After the removal of accumulated substances by washing, the
164
surface area of hydrochar increased significantly, especially after the first three THF washings,
165
and only negligible changes occurred thereafter. This phenomenon agrees well with the weight
166
loss behavior of hydrochar, which is further confirmed by the positive correlation between the
167
washing efficiency of hydrochar (loss ratio of hydrochar mass) and the increased surface area of
168
hydrochar samples (Figure 1a).
169
As shown in Table 1, C content experienced a lower loss ratio than H content while at a higher
170
loss ratio than O content. Such loss ratio difference resulted in decreased H/C and increased O/C
171
atomic ratios for the THF-washed hydrochar. Thus, THF washing increased the aromaticity and
9 ACS Paragon Plus Environment
Page 11 of 29
Environmental Science & Technology
172
polarity of the hydrochar. The derived Van Krevelen diagram suggests that THF-based hydrochar
173
washing was mainly followed a demethanation trend (Figure 1b), attributable to higher C and H
174
molar ratios than that of O in the extracted bio-oil, such as aromatic phenol, aliphatic acid and
175
aliphatic alcohols. This was partially confirmed from the results of GC-MS analysis.
176
It has been well documented that increased aromaticity (decreased H/C atomic ratio) can
177
enhance the stability of carbonaceous material.23 As shown in Figure 1c and Table S2, THF
178
washing introduced great differences in TG characteristics of hydrochar. As expected, THF
179
washing enhanced the stability of hydrochar, as indicated by the increases of R50 (novel
180
recalcitrance index) and W700 (residual weight at 700 °C in TG curves)24, 25. The detailed R50
181
calculation could be seen in the note of Table S2. In addition, these two indices are strongly
182
negatively correlated with the H/C ratio of hydrochar (R2 = 0.97 and 0.96, respectively) (Figure
183
S3), suggested that the aromaticity of hydrochar also improved by THF washing. A low H/C
184
value represented the high aromaticity. The pyrolysis behavior of hydrochar can be divided into
185
the following two regions. The first region of the differential thermogravimetry (DTG) curves
186
was predominantly due to the decomposition of small molecular compounds (including GC-MS
187
detectable matters) with low peak temperatures (Tmax1). After THF washing, the weight loss of
188
hydrochar in this region decreased substantially (Table S2). It is clear that few compounds such
189
as aromatic phenol, aliphatic acid, and aliphatic alcohol could be detected in the GC-MS analysis
190
after the second THF washing (Table S1), which is well corresponded with the Tmax1
191
characteristics in DTG profiles of this region (Figure 1c and Table S2). The second region was
10 ACS Paragon Plus Environment
Environmental Science & Technology
192
mainly affected by the loss of aromatic species.26 The increase in peak temperature (Tmax2) was
193
easily detected along with washing times and negatively controlled by the H/C ratio of hydrochar
194
(Figure S3), which also implies the increase in thermal stability of the hydrochar.
Page 12 of 29
195
FTIR, NMR, and XPS spectra were used to further confirm the changes of hydrochar
196
characteristics after different THF washing times. As shown in Figure 1d, the band intensities at
197
~ 2920 and 2845 cm-1 (aliphatic C-H stretching) were obviously weakened with the increase of
198
washing times, further confirming the demethanation trend of hydrochar during the washing
199
process. Due to the removal of O-containing matters (i.e., aromatic phenol, aliphatic acid, and
200
aliphatic alcohol), the band intensities at ~ 1705 cm-1 (C=O stretching), ~ 1204 and 1120 cm-1
201
(C-O stretching) also decreased with the increase of washing times. Smaller changes were
202
observed for aromatic C=C (~ 1600 cm-1), lignin C=C (~ 1505 cm-1), and aromatic ring (~ 1450
203
cm-1) FTIR bands,27,
204
solvent-washing process.
28
indicating that aromatic compounds were more stable during the
205
As shown in the 13C NMR spectra of hydrochar, the relative proportion of alkyl C (0-43 ppm)
206
decreased with the increase of washing times (Figure 1e and Table S3), resulting in the
207
demethanation trend and decreased H/C ratio. These results were confirmed by the positive
208
correlation between the relative ratio of alkyl C and H/C atomic ratio (Figure S4). The removal
209
of alkyl C resulted in a conspicuous increase of aromatic C (110-145 ppm) and aromatic C-O
210
(145-160 ppm), further reflecting an increased hydrochar aromaticity and thermal stability after
211
THF washing. Due to a higher retention of O element in the THF-washed hydrochar samples, the
11 ACS Paragon Plus Environment
Page 13 of 29
Environmental Science & Technology
212
O-containing groups like O-CH3 (43-60 ppm), O-alkyl C (60-90 ppm), carboxyl/ester (160-190
213
ppm) and carbonyl (190-220 ppm) became more evident.
214
As presented in Figure 1f, four representative peaks were observed in the hydrochar, which
215
could be attributed to C1 (284.6 eV for C-H/C-C, aliphatic/aromatic carbon groups), C2 (285.6
216
eV for C-O, hydroxyl/ether groups), C3 (287 eV for C=O, carbonyl/quinone groups), C4 (289 eV
217
for O-C=O, carboxylic/ester/lactone groups).29 The XPS results reveal that C-O was the main
218
O-containing functional group on the hydrochar surface, corresponding with the higher
219
intensities of aromatic C-O, O-CH3 and O-alkyl than the carboxyl and carbonyl groups in the
220
NMR spectra. After THF washing, the relative intensity of C-C/C-H obviously decreased.
221
Accordingly, the relative intensity of C-O and C=O obviously increased (Table S4), further
222
indicating that demethanation dominated the THF washing of hydrochar.
223
Overall, the mass loss of hydrochar was mainly occurred during the first three times of THF
224
washing, and demethanation reflected the behavior of hydrochar during THF washings.
225
Accordingly, THF washing could gradually bring up O content, aromaticity, and stability while
226
bring down C and H contents, as well as the H/C atomic ratio of hydrochar. This is accompanied
227
by decreased intensities of aliphatic C-H stretching in the FTIR spectra, alkyl C in the 13C NMR
228
spectra and C1s for C-C/C-H in the XPS spectra.
229
Dynamic Changes in Hydrochar Characteristics Under Different Solvents
230
As shown in Table S5, the solvents used for hydrochar washing showed significant effects on
231
the loss ratio of hydrochar. The loss ratio of hydrochar using four solvents was in a descending 12 ACS Paragon Plus Environment
Environmental Science & Technology
Page 14 of 29
232
order of PE < DLM < AT < THF. THF exhibited the highest performance during the washing
233
process, possibly due to its strong solubility in bio-oil.13 GC-MS analysis further confirmed this
234
result (Figure S5 and Table S6). Obviously, THF possessed the highest capability for extracting
235
the GC-MS detectable matters such as aromatic phenol, aliphatic acid, and aliphatic alcohol.
236
Accordingly, the surface area of washed hydrochar was increased and positively controlled by
237
their washing efficiency (Figure S6), revealing the pore blockage effect of bio-oil composition on
238
the porosity of hydrochar. SEM spectra indicate that the hydrochar obviously became rough and
239
some pores could be observed after AT and THF washings, due to more bio-oil removal (Figure
240
2).
241
As shown in Figure S7, the loss ratio of hydrochar mass is a good indicator for forecasting C,
242
H and O contents and H/C and O/C atomic ratios. A decrease in H/C atomic ratio and an increase
243
in O/C atomic ratio were observed in all solvent-washed hydrochars, probably attributable to the
244
same descending order of loss ratio in elemental composition as H content > C content > O
245
content (Table S5). This might result in the tendency of demethanation of hydrochar washed by
246
PE, DLM, AT, and THF, as confirmed by the typical Van Krevelen diagram (Figure S8).
247
FTIR, NMR and XPS were also used to examine the characteristics of hydrochar washed with
248
the different solvents. Due to PE’s weak extraction ability of aromatic phenol, aliphatic acid and
249
aliphatic alcohol (confirmed in the GC-MS spectrum in Figure S5), PE-washed hydrochar
250
retained stronger signals for C-H stretching (~ 2920 and 2845 cm-1), C=O stretching (~ 1705
251
cm-1), and C-O stretching (~ 1204 and 1120 cm-1) in Figure S9. In contrast, THF-washed
13 ACS Paragon Plus Environment
Page 15 of 29
Environmental Science & Technology
252
hydrochar exhibited relatively weak signals in the aforementioned spectrum bands due to its
253
strong extraction ability of bio-oil compositions. Similar trends could also be discerned in the 13C
254
NMR (Figure S10 and Table S7) like decreased alkyl C (0 - 43 ppm), and in the C1s XPS spectra
255
like the decreased C-C/C-H intensity (Figure S11 and Table S8). These three technical
256
characteristics confirmed the effectiveness of PE, DLM, AT and THF in hydrochar washing, and
257
they provide convincing evidence that solvent-washing of hydrochar is regulated by
258
demethanation process.
259
Modification of Washed Hydrochar by Simultaneous Activation and Magnetization
260
Different solvent-washed hydrochar samples were further modified for MC preparation. The
261
yields (based on the same weight of washed hydrochar) and elemental composition of MCs
262
remained very similar with respect to the changes in hydrochar properties (Table S9). Little
263
difference in the yields of MCs indicates that the aromaticity of solvent-washed hydrochar is not
264
a key factor in controlling the properties of the derived MCs. But, MCs yields based on the
265
weight of raw un-washed hydrochar (H-WT) indicated the adverse effect of solvent washing, due
266
to the removal of bio-oil in the hydrochar washing process.
267
As shown clearly, solvent washing of hydrochar enhanced the BET surface area and micropore
268
surface area of the derived MCs (Table 2 and Figure S12), which was further confirmed by a
269
positive correlation between the surface area of hydrochar and that of MC (R2 = 0.95, Figure 3a).
270
This result suggests that a hydrochar with higher porosity provides more tunnels for storing
271
ZnCl2 (activator) in the impregnation stage, thereby promoting the activation of hydrochar and 14 ACS Paragon Plus Environment
Environmental Science & Technology
Page 16 of 29
272
producing a MC material with high surface area. The main micropore sizes of samples
273
(concluded from CO2 adsorption isotherms, Figure S12 c) were 0.36, 0.48, 0.52, 0.60 and 0.82
274
nm (Figure 3b). It also should be pointed out that the micropore sizes of MC were not influenced
275
by the solvent washing of hydrochar. Noticeably, solvent washing of hydrochar reduced the
276
microporosity of the derived MCs (indicated by Smic/SBET and Vmic/Vt in Table 2), suggesting that
277
the excess entrance of ZnCl2 into the pores probably caused the collapse of the MCs’ pore walls.
278
Therefore, hydrochar pores play a dual role in the porosity of hydrochar-derived MCs.
279
BPA has toxicity and is widely detected in the environment, so it was selected to examine the
280
adsorption capacity of MCs.30 The adsorption data were fitted to the Langmuir model for the
281
calculation of maximum adsorption capacity (Note of Figure S13). Due to the pronounced
282
pore-filling effect, the MC-THF possessed the highest BPA adsorption capacity (Figure S13a).19
283
This result was further confirmed by the positive correlations between the surface area, the
284
micropore surface area of MCs, and the qm of BPA adsorption (Figure S13b and S13c). BPA
285
possesses a three-dimensional size of 0.383 × 0.587 × 1.068 nm,31 suggesting that MC
286
micropores are better suited for filling BPA molecules. Further confirmation was provided by the
287
stronger correlation between the micropore surface area of MCs and qm than that between the
288
surface area of MCs and qm (Figure S13b and S13c).
289
XRD patterns showed that Fe2O3 (JCPDS 39-1346), Fe3O4 (JCPDS 65-3107), and ZnFe2O4
290
(JCPDS 22-1012) would be the main Fe species (Figure S14). However, XRD analysis cannot
291
effectively differentiate these three species, because they have similar characteristic peaks.
15 ACS Paragon Plus Environment
Page 17 of 29
Environmental Science & Technology
292
Therefore, Mössbauer spectroscopy was used to further distinguish the Fe species. As shown in
293
Figure 4, Fe3O4 and ZnFe2O4 were detected as the main Fe species. Based on the following
294
equations (3-5), Fe3O4 is produced from a reduction reaction between Fe2O3 and carbon matrix32,
295
33
296
reaction process between ZnO (acidic oxide) and Fe2O3 (basic oxide).22, 34
; ZnFe2O4 may be formed on the surface of carbon matrix via an aggregation thermochemical
Fe3+ → Fe(OH)3 → FeOOH → Fe2 O3 297
T ≤ 400 o C
(3)
3Fe2 O3 + C → 2Fe3 O 4 + CO
T ≥ 500 o C
(4)
Fe2 O3 + ZnO → ZnFe 2 O4
T ≥ 500 o C
(5)
298
It has been noticed that all MCs derived from organic solvent-washed hydrochar (PE, DLM,
299
AT, and THF) possess a lower ZnFe2O4 mole fraction. Before the activation stage of these
300
samples, more ZnCl2 (a precursor of ZnO) activator would be dispersed into the pores of
301
hydrochar due to its higher porosity (Figure S6). Correspondingly, more ZnO (produced from the
302
decomposition of ZnCl2) is loaded into the pores of MC. Undoubtedly, this will inhibit the
303
evolution of surface reaction between Fe2O3 and ZnO and therefore decrease the content of
304
ZnFe2O4 in MCs derived from organic solvent-washed hydrochar.
305
As shown in Figure S15a, the MCs had similar saturation magnetization values, ranging from
306
19.1 to 25.5 emu/g. However, the solubility of Fe species increased linearly with the increase of
307
MC porosity (Figure S15b), suggesting its decreased acid resistance in the MC derived from
308
organic solvent-washed hydrochar.
309
Overall, because solvent washing promotes porosity formation in hydrochar, the surface area
310
and adsorption capacities of MCs increased accordingly. For the same reason, MCs’
16 ACS Paragon Plus Environment
Environmental Science & Technology
311
microporosity, ZnFe2O4 content, and acid resistance were found decreased.
312
Environmental Implications
Page 18 of 29
313
As discussed above, solvent-washing process significantly altered hydrochar properties in
314
terms of element content, group type and porosity. As elaborated below, the findings from the
315
hydrochar washing mechanism have significant implications for the applications of hydrochar
316
especially in environmental remediation field. Firstly, changes in O-containing group content of
317
will affect the ability of hydrochar on reducing metal ions and organic matters, because carbonyl
318
groups are thought to be responsible for reduction capability of carbonaceous material.35 The
319
content change will also influence the catalytic reactivity of hydrochar in oxidation of organic
320
matters, because O-containing groups act as an efficient activator in persulfate-based
321
oxidations.36 Also, the content change will affect the adsorption of organic matters onto the
322
hydrochar surface through the π-π electron-donor-acceptor interaction.3, 17, 37 In addition, the
323
enhanced porosity of washed hydrochar will likely contribute to the adsorption of organic
324
matters onto hydrochar surface via the pore-filling effect.
325
Lastly, due to the removal of bio-oil, organic solvent washing process could reduce the release
326
of dissolved organic matter from hydrochar and improve the carbon sequestration potential of
327
hydrochar. Thus, organic solvent washing procedure of hydrochar has important implications for
328
the global carbon cycle.
329
ACKNOWLEDGEMENTS
330
This research was funded by the National Key Technology Support Program (No. 17 ACS Paragon Plus Environment
Page 19 of 29
Environmental Science & Technology
331
2015BAD15B06), the National Natural Science Foundation of China (No. 21407027, 21577025),
332
and the International Postdoctoral Exchange Fellowship Program of China Supported by Fudan
333
University.
334
Supporting Information
335
The pseudo-second-order models, GC-MS spectrum, surface area, Van Krevelen diagram,
336
FTIR spectra, 13C NMR spectra, C 1s XPS spectra, N2 adsorption-desorption isotherms, pore size
337
distribution, adsorption isotherms of BPA onto MCs, XRD patterns, hysteresis loop, correlation
338
analysis of related samples are presented in Figure S1 - S15.
339
GC-MS analysis of bio-oil, BET surface area, TG characteristics, loss characteristics, element
340
compositions, C1s analysis for XPS spectra, C-containing functional groups analysis for NMR
341
spectra of hydrochar, and yields, elemental compositions of MCs are presented in Table S1 - S9.
342
This material is available free of charge via the Internet at http://pubs.acs.org.
343
18 ACS Paragon Plus Environment
Environmental Science & Technology
Page 20 of 29
344
REFERENCES
345
(1) Huggins, T. M.; Haeger, A.; Biffinger, J. C.; Ren, Z. J. Granular biochar compared with activated carbon
346
for wastewater treatment and resource recovery. Water Res. 2016, 94, 225-232.
347
(2) Huggins, T. M.; Pietron, J. J.; Wang, H.; Ren, Z. J.; Biffinger, J. C. Graphitic biochar as a cathode
348
electrocatalyst support for microbial fuel cells. Bioresour. Technol. 2015, 195, 147-153.
349
(3) Han, L.; Ro, K. S.; Sun, K.; Sun, H.; Wang, Z.; Libra, J. A.; Xing, B. New evidence for high sorption
350
capacity of hydrochar for hydrophobic organic pollutants. Environ. Sci. Technol. 2016, 50, 13274-13282.
351
(4) Berge, N. D.; Ro, K. S.; Mao, J.; Flora, J. R. V.; Chappell, M. A.; Bae, S. Hydrothermal carbonization of
352
municipal waste streams. Environ. Sci. Technol. 2011, 45, 5696-5703.
353
(5) Heilmann, S. M.; Molde, J. S.; Timler, J. G.; Wood, B. M.; Mikula, A. L.; Vozhdayev, G. V.; Colosky, E. C.;
354
Spokas, K. A.; Valentas, K. J. Phosphorus reclamation through hydrothermal carbonization of animal manures.
355
Environ. Sci. Technol. 2014, 48, 10323-10329.
356
(6) He, C.; Wang, K.; Yang, Y.; Amaniampong, P. N.; Wang, J. Y. Effective nitrogen removal and recovery
357
from dewatered sewage sludge using a novel integrated system of accelerated hydrothermal deamination and
358
air stripping. Environ. Sci. Technol. 2015, 49, 6872-6880.
359
(7) Mumme, J.; Titirici, M.-M.; Pfeiffer, A.; Lüder, U.; Reza, M. T.; Mašek, O. e. Hydrothermal carbonization
360
of digestate in the presence of zeolite: process efficiency and composite properties. ACS Sustainable Chem.
361
Eng. 2015, 3, 2967-2974.
362
(8) Zhu, X.; Liu, Y.; Zhou, C.; Zhang, S.; Chen, J. Novel and high-performance magnetic carbon composite
363
prepared from waste hydrochar for dye removal. ACS Sustainable Chem. Eng. 2014, 2, 969-977.
364
(9) Zhu, X.; Liu, Y.; Qian, F.; Zhou, C.; Zhang, S.; Chen, J. Role of hydrochar properties on the porosity of
365
hydrochar-based porous carbon for their sustainable application. ACS Sustainable Chem. Eng. 2015, 3,
366
833-840.
367
(10) Xue, Y.; Gao, B.; Yao, Y.; Inyang, M.; Zhang, M.; Zimmerman, A. R.; Ro, K. S. Hydrogen peroxide
368
modification enhances the ability of biochar (hydrochar) produced from hydrothermal carbonization of peanut
369
hull to remove aqueous heavy metals: batch and column tests. Chem. Eng. J. 2012, 200, 673-680.
370
(11) Sun, Y.; Gao, B.; Yao, Y.; Fang, J.; Zhang, M.; Zhou, Y.; Chen, H.; Yang, L. Effects of feedstock type,
19 ACS Paragon Plus Environment
Page 21 of 29
Environmental Science & Technology
371
production method, and pyrolysis temperature on biochar and hydrochar properties. Chem. Eng. J. 2014, 240,
372
574-578.
373
(12) He, C.; Zhao, J.; Yang, Y.; Wang, J. Y. Multiscale characteristics dynamics of hydrochar from
374
hydrothermal conversion of sewage sludge under sub-and near-critical water. Bioresour. Technol. 2016, 211,
375
486-493.
376
(13) Yang, X.; Lyu, H.; Chen, K.; Zhu, X.; Zhang, S.; Chen, J. Selective extraction of bio-oil from
377
hydrothermal liquefaction of salix psammophila by organic solvents with different polarities through multistep
378
extraction separation. BioResources 2014, 9, 5219-5233.
379
(14) Li, C.; Yang, X.; Zhang, Z.; Zhou, D.; Zhang, L.; Zhang, S.; Chen, J. Hydrothermal liquefaction of desert
380
shrub Salix psammophila to high value-added chemicals and hydrochar with recycled processing water.
381
BioResources 2013, 8, 2981-2997.
382
(15) Deng, H.; Meredith, W.; Uguna, C. N.; Snape, C. E. Impact of solvent type and condition on biomass
383
liquefaction to produce heavy oils in high yield with low oxygen contents. J. Anal. Appl. Pyrol. 2015, 113,
384
340-348.
385
(16) Elliott, D. C.; Biller, P.; Ross, A. B.; Schmidt, A. J.; Jones, S. B. Hydrothermal liquefaction of biomass:
386
Developments from batch to continuous process. Bioresour. Technol. 2015, 178, 147-156.
387
(17) Jing, X. R.; Wang, Y. Y.; Liu, W. J.; Wang, Y. K.; Jiang, H. Enhanced adsorption performance of
388
tetracycline in aqueous solutions by methanol-modified biochar. Chem. Eng. J. 2014, 248, 168-174.
389
(18) Hao, W.; Björkman, E.; Yun, Y.; Lilliestråle, M.; Hedin, N. Iron oxide nanoparticles embedded in
390
activated carbons prepared from hydrothermally treated waste biomass. ChemSusChem 2014, 7, 875-882.
391
(19) Zhu, X.; Qian, F.; Liu, Y.; Zhang, S.; Chen, J. Environmental performances of hydrochar-derived
392
magnetic carbon composite affected by its carbonaceous precursor. RSC Adv. 2015, 5, 60713-60722.
393
(20) Falco, C.; Marco-Lozar, J. P.; Salinas-Torres, D.; Morallon, E.; Cazorla-Amorós, D.; Titirici, M.-M.;
394
Lozano-Castelló, D. Tailoring the porosity of chemically activated hydrothermal carbons: influence of the
395
precursor and hydrothermal carbonization temperature. Carbon 2013, 62, 346-355.
396
(21) Zhu, X.; Liu, Y.; Luo, G.; Qian, F.; Zhang, S.; Chen, J. Facile fabrication of magnetic carbon composites
397
from hydrochar via simultaneous activation and magnetization for triclosan adsorption. Environ. Sci. Technol.
20 ACS Paragon Plus Environment
Environmental Science & Technology
Page 22 of 29
398
2014, 48, 5840-5848.
399
(22) Zhu, X.; Qian, F.; Liu, Y.; Matera, D.; Wu, G.; Zhang, S.; Chen, J. Controllable synthesis of magnetic
400
carbon composites with high porosity and strong acid resistance from hydrochar for efficient removal of
401
organic pollutants: an overlooked influence. Carbon 2016, 99, 338-347.
402
(23) Zhu, X.; Liu, Y.; Qian, F.; Zhang, S.; Chen, J. Investigation on the physical and chemical properties of
403
hydrochar and its derived pyrolysis char for their potential application: influence of hydrothermal
404
carbonization conditions. Energ. Fuel. 2015, 29, 5222-5230.
405
(24) Harvey, O. R.; Kuo, L.-J.; Zimmerman, A. R.; Louchouarn, P.; Amonette, J. E.; Herbert, B. E. An
406
index-based approach to assessing recalcitrance and soil carbon sequestration potential of engineered black
407
carbons (biochars). Environ. Sci. Technol. 2012, 46, 1415-1421.
408
(25) Tian, Y.; Zhang, J.; Zuo, W.; Chen, L.; Cui, Y.; Tan, T. Nitrogen conversion in relation to NH3 and HCN
409
during microwave pyrolysis of sewage sludge. Environ. Sci. Technol. 2013, 47, 3498-3505.
410
(26) Tian, K.; Liu, W. J.; Qian, T. T.; Jiang, H.; Yu, H. Q. Investigation on the evolution of N-containing
411
organic compounds during pyrolysis of sewage sludge. Environ. Sci. Technol. 2014, 48, 10888-10896.
412
(27) Li, F.; Cao, X.; Zhao, L.; Wang, J.; Ding, Z. Effects of mineral additives on biochar formation: carbon
413
retention, stability, and properties. Environ. Sci. Technol. 2014, 48, 11211-11217.
414
(28) Xiao, X.; Chen, B.; Zhu, L. Transformation, morphology, and dissolution of silicon and carbon in rice
415
straw-derived biochars under different pyrolytic temperatures. Environ. Sci. Technol. 2014, 48, 3411-3419.
416
(29) Sevilla, M.; Fuertes, A. B. Chemical and structural properties of carbonaceous products obtained by
417
hydrothermal carbonization of saccharides. Chem. Eur. J. 2009, 15, 4195-4203.
418
(30) Hu, L.; Fong, C. C.; Zhang, X.; Chan, L. L.; Lam, P. K. S.; Chu, P. K.; Wong, K. Y.; Yang, M. Au
419
nanoparticles decorated TiO2 nanotube arrays as a recyclable sensor for photoenhanced electrochemical
420
detection of bisphenol A. Environ. Sci. Technol. 2016, 50, 4430-4438.
421
(31) Nghiem, L. D.; Vogel, D.; Khan, S. Characterising humic acid fouling of nanofiltration membranes using
422
bisphenol A as a molecular indicator. Water Res. 2008, 42, 4049-4058.
423
(32) Liu, W. J.; Tian, K.; Jiang, H.; Yu, H. Q. Facile synthesis of highly efficient and recyclable magnetic solid
424
acid from biomass waste. Sci. Rep. 2013, 2419, 1-7.
21 ACS Paragon Plus Environment
Page 23 of 29
Environmental Science & Technology
425
(33) Yang, J.; Zhao, Y.; Ma, S.; Zhu, B.; Zhang, J.; Zheng, C. Mercury removal by magnetic biochar derived
426
from simultaneous activation and magnetization of sawdust. Environ. Sci. Technol. 2016, 50, 12040-12047.
427
(34) Sui, J.; Zhang, C.; Hong, D.; Li, J.; Cheng, Q.; Li, Z.; Cai, W. Facile synthesis of MWCNT-ZnFe2O4
428
nanocomposites as anode materials for lithium ion batteries. J. Mater. Chem. 2012, 22, 13674-13681.
429
(35) Xu, Y.; Lou, Z.; Yi, P.; Chen, J.; Ma, X.; Wang, Y.; Li, M.; Chen, W.; Liu, Q.; Zhou, J. Improving abiotic
430
reducing ability of hydrothermal biochar by low temperature oxidation under air. Bioresour. Technol. 2014,
431
172, 212-218.
432
(36) Fang, G.; Liu, C.; Gao, J.; Dionysiou, D. D.; Zhou, D. Manipulation of persistent free radicals in biochar
433
to activate persulfate for contaminant degradation. Environ. Sci. Technol. 2015, 49, 5645-5653.
434
(37) Qi, X.; Li, L.; Tan, T.; Chen, W.; Smith, R. L. Adsorption of 1-butyl-3-methylimidazolium chloride ionic
435
liquid by functional carbon microspheres from hydrothermal carbonization of cellulose. Environ. Sci. Technol.
436
2013, 47, 2792-2798.
437 438
22 ACS Paragon Plus Environment
a y=0.31x-6.55 R2=0.99, p < 0.01
12
H-3
1.1
H-4 H-2
9 6 3
H-1
o
0.0 H-1 H-0 H-5 Tmax1
60
-0.2
50
200 400 600 Temperature (oC)
40
200
H-5 H-4
H-3
H-4
H-3 H-2 Tmax2
-0.3
e
H-4
aromatic C-O C=O
H-5 H-2 H-1 H-0
800
400 600 Temperature (oC)
1.5
1.0
1.0
Decarboxylation
0.25 0.50 0.75 O/C (atomic ratio)
d H-5
1.00
aromatic aromatic ring C-O C=C
aliphatic C-H
H-4 H-3 H-2 H-1 H-0 C=O
4000
lignin, C=C
3000 2000 1000 Wavenumber (cm-1) H-0
f
H-5 0.29
0.28
0.5
800
aromatic C O-alkyl C O-CH3 alkyl C
H-3
H-4
0.27
Transmittance (arbitrary units)
80 Derivate Weight (%/ C)
Weight (%)
90
-0.1
H-0 H-1
0.0 0.00
30 40 50 60 70 Washing efficiency of hydrochar by THF (%) c 100
70
b H-2
D eh yd ra D tio em n et ha na tio n
15
Page 24 of 29
2.0
H-5 H/C (atomic ratio)
BET surface area increasement (m2/g)
Environmental Science & Technology
H-1
H-2
C1s (1)
C1s (1)
C1s (1)
C1s (2)
C1s (2)
C1s (2) C1s (3)
C1s (3) C1s (4)
COO
290
288
C1s (3)
C1s (4) 286
282 290
284
C1s (4)
288
286
284
282 290
288
286
284
282
H-3 H-3
H-4
C1s (1)
H-2
C1s (2)
H-1
150 100 50 Chemical shift (ppm)
288
C1s (2) C1s (3)
C1s (3)
C1s (4)
C1s (4)
C1s (4) 290
200
C1s (1)
C1s (2)
C1s (3)
H-0
H-5
C1s (1)
286
284
282 290
0
288
286
284
282 290
288
286
284
Binding Energy (eV)
Figure 1 (a) Correlation between washing efficiency of hydrochar by THF and increase in BET surface area, dynamics in (b) H/C and O/C atomic ratios using Van Krevelen diagram, (c) TG and DTG profiles, (d) FTIR spectra, (e)
13
C NMR spectra, and (f) C 1s XPS spectra for
hydrochar after different THF washing times.
23 ACS Paragon Plus Environment
282
Page 25 of 29
Environmental Science & Technology
H-WT
H-PE
H-AT
H-THF
Figure 2 SEM spectra for the hydrochar washed by different solvents.
24 ACS Paragon Plus Environment
Environmental Science & Technology
1.2 b
920 880 MC-PE
MC-DLM
0.52 nm
y=27.11x+834.8 R2=0.95 p < 0.01 MC-AT
960
1.0 0.8 0.6 0.4
MC-WT MC-AT MC-DLM MC-THF MC-PE
0.48 nm
0.60 nm 0.82 nm
0.36 nm
MC-THF dV(D) (cm3/nm/g)
Surface area of MC (m2/g)
1000 a
Page 26 of 29
0.2 840 0.0
MC-WT 0
1 2 3 4 5 Surface area of Hydrochar (m2/g)
6
0.4
0.6 0.8 1.0 1.2 Pore diameter (nm)
1.4
Figure 3 (a) Positive correlation between surface area of hydrochars and magnetic carbon composites derived from hydrochar being washed by different solvents, (b) narrow micropore size distribution from CO2 adsorption isotherms for magnetic carbon composites derived from hydrochar being washed by different solvents.
25 ACS Paragon Plus Environment
Page 27 of 29
Environmental Science & Technology
ZnFe2O4
Relative Transmission
ZnFe2O4
ZnFe2O4 Fe3O4
Fe3O4
Fe3O4
MC-WT
MC-DLM
MC-PE -10
Relative Transmission
0 5 Velocity (mm/s)
1.0
10
MC-WT
1.5
Fe3O4
Fe3O4
ZnFe2O4/Fe3O4 (mole ratio)
ZnFe2O4
ZnFe2O4
-5
MC-AT MC-DLM MC-PE MC-THF
0.5
MC-THF
MC-AT -10
-5
0 5 Velocity (mm/s)
10
0.0
-10
-5
0 5 Velocity (mm/s)
10
Figure 4 Room-temperature Mössbauer spectra and mole ratio of ZnFe2O4 and Fe3O4 for magnetic carbon composites derived from hydrochar after being washed by different solvents.
26 ACS Paragon Plus Environment
Environmental Science & Technology
Page 28 of 29
Table 1 Loss Characteristics and Elemental Compositions of Hydrochar Samples Washed by THF for 0 ~ 5 Times Loss ratio (%) Sample
a
Ash
C
O
H
N H/C
O/C
0.573
1.59
0.192
8.37
0.619
1.47
0.209
23.6
6.12
0.696
1.12
0.269
63.4
24.7
5.78
0.684
1.09
0.292
5.45
64.9
23.0
5.95
0.697
1.10
0.266
4.46
64.4
25.2
5.21
0.683 0.970 0.294
Massa
Cb
Oc
Hd
Ne
(%)
(%)
(%)
(%)
(%)
H-0
-
-
-
-
-
2.92
69.5
17.8
9.20
H-1
31.7
32.6
26.8
37.8
26.2
3.42
68.5
19.1
H-2
58.7
60.9
45.3
72.5
49.8
3.81
65.8
H-3
66.2
69.1
53.2
78.7
59.6
5.42
H-4
70.9
72.8
62.4
81.1
64.6
H-5
73.4
75.3
62.3
84.9
68.3
Loss ratio of mass = (1-hydrochar weight after washing/raw hydrochar weight)*100%, b loss
ratio of C content = (1-C content of hydrochar after washing/raw C content of hydrochar)*100%, c
loss ratio of O content = (1-O content of hydrochar after washing/raw O content of
hydrochar)*100%, d loss ratio of H content = (1-H content of hydrochar after washing/raw H content of hydrochar)*100%,
e
loss ratio of N content = (1-N content of hydrochar after
washing/raw N content of hydrochar)*100%.
27 ACS Paragon Plus Environment
Page 29 of 29
Environmental Science & Technology
Table 2 Surface Area and Pore Volume Parameters for Magnetic Carbon Composites Derived from Different Solvents Washed Hydrochar SBET
Smic
Smic/SBET
Vt
Vmic
Vmic/Vt
Smic
Vmic
(m2/g)a
(m2/g)a
(%)
(cm3/g)a
(cm3/g)a
(%)a
MC-WT
821.6
785.4
95.6
0.427
0.389
91.0
675.9
0.226
MC-PE
858.9
802.3
93.4
0.454
0.396
87.2
765.0
0.255
MC-DLM
897.3
843.8
94.0
0.471
0.413
87.7
862.9
0.295
MC-AT
912.5
840.5
92.1
0.522
0.417
79.9
777.5
0.262
MC-THF
1001
940.5
94.0
0.559
0.466
83.3
864.5
0.299
Sample
a
Porosity parameters concluded from N2 adsorption isotherms.
b
Porosity parameters concluded from CO2 adsorption isotherms.
28 ACS Paragon Plus Environment
(m2/g)b (cm3/g)b