You are on page 1of 7

Article

Cite This: J. Phys. Chem. C XXXX, XXX, XXX−XXX pubs.acs.org/JPCC

High CO2 Sensitivity and Reversibility on Nitrogen-Containing


Polymer by Remarkable CO2 Adsorption on Nitrogen Sites
Sharif Md Khan,† Hiroki Kitayama,† Yasuhiro Yamada,‡ Syun Gohda,§ Hironobu Ono,§
Daichi Umeda,‡ Kouki Abe,‡ Kenji Hata,∥ and Tomonori Ohba*,†

Graduate School of Science and ‡Graduate School of Engineering, Chiba University, 1-33 Yayoi, Inage, Chiba 263-8522, Japan
§
Nippon Shokubai Co., Ltd., 5-8 Nishiotabi, Suita, Osaka, 564-0034, Japan

National Institute of Advanced Industrial Science and Technology (AIST), 1-1-1 Higashi Tsukuba, Ibaraki 305-8565, Japan
*
Downloaded via INDIAN ASSOCIATION CULTIVATION SCI on October 10, 2018 at 06:45:24 (UTC).

S Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: CO2 removal and sensing, especially by amine groups, are necessary to prevent
global warming. However, understanding the dynamic mechanism of CO2 capture on nitrogen
sites remains challenging. We fabricated a nitrogen-doped polymer and compared the
mechanism of CO2 adsorption with that of a similar polymer without nitrogen atoms by
evaluating the CO2 adsorption at 243, 273, and 303 K; electrical resistance changes due to
CO2 adsorption at 297 K; and molecular dynamics simulations. The CO2 adsorption of the
nitrogen-doped polymer was significantly improved owing to the negative partial charges on
nitrogen sites. In addition, the electrical response of the nitrogen-doped polymer to CO2
showed adequate reversibility. These results originated from the stronger adsorption of CO2
on nitrogen sites than on ring and vinyl sites. Our findings contribute to the understanding of
CO2−nitrogen attraction and will be beneficial for the development of efficient CO2 capture
and sensing.

1. INTRODUCTION recently attracted great interest because of their low density,


Increasing atmospheric carbon dioxide is considered the major large surface area, and high porosity.29−33
Doping of heteroatoms such as nitrogen atoms into porous
reason for climate change and global warming.1,2 The
polymers and porous carbons can markedly increase the
concentration of CO2 in the atmosphere has significantly
amount of adsorbed CO2 and CO2/N2 selectivity.34−36 Chen
increased, reaching 400 ppm in 2013.3 To reduce the
et al. reported the high CO2 capture performance of
temperature increase to 2.48−2.88 K by increase of CO2 microporous polycarbazole of 4.8 mmol g−1 at 273 K and
concentration, it is essential to maintain the atmospheric 0.1 MPa.37 Hug et al. synthesized a nitrogen-rich subclass of
concentration of CO2 below 450 ppm. To achieve this porous polymers with significant CO2 capability up to 4.28
objective, CO2 emissions must be reduced by 30%−60% by mmol g−1.38 In addition, a benzimidazole-linked porous
2050.4,5 Primary energy sources based on fossil fuels (coal, organic polymer with nitrogen sites was shown to exhibit
petroleum, and natural gas) make up ∼86% of all energy excellent CO2 capability of 5.3 mmol g−1 at 273 K and 0.1
sources, and it will be difficult to restrict the atmospheric CO2 MPa.39,40 Li et al. prepared nitrogen-enriched mesoporous
concentration in the upcoming decades as fossil fuels remain carbon spheres using a nitrogen-containing precursor, and their
the most useful energy source, especially in developing CO2 capacity reached 2.9 mmol g−1 at 0.1 MPa and 298 K.41
countries. 6 To reduce the CO 2 concentration in the Despite the effectiveness of using a nitrogen-containing
atmosphere, carbon capture and storage as well as carbon precursor and nitrogen-enriched adsorbent materials for CO2
capture and utilization are being explored worldwide. capture, the mechanism of CO2 capture by nitrogen donation
Adsorption and membrane separation have been considered remains unclear. The enhancement of CO2 capture on
a suitable technology for CO2 capture owing to the low energy nitrogen-containing surfaces is attributed to the acid−base
requirements.7 Porous materials such as porous carbon,8−13 interactions between nitrogen-containing basic functional
metal−organic frameworks,14,15 zeolites,16−19 porous organic groups and acidic CO2 molecules.42 Xing et al. proposed
polymers,20 nanoporous dipeptide-based materials,21 amine- that N-doped activated carbons adsorbed CO2 through
modified silicas,22,23 and others24−28 have been available as enhancement of hydrogen bonding between the carbon surface
efficient materials for CO2 removal. However, porous carbons and CO2.43 In contrast, Sevilla et al. suggested that nitrogen
have limited CO2/N2 selectivity, and zeolites are inefficient at
capturing CO2 from flue gases with water vapor because of Received: July 31, 2018
their strong hydrophilic nature.6 Polymer-based adsorbent Revised: September 14, 2018
materials with high thermal and hydrothermal stability have Published: September 24, 2018

© XXXX American Chemical Society A DOI: 10.1021/acs.jpcc.8b07420


J. Phys. Chem. C XXXX, XXX, XXX−XXX
The Journal of Physical Chemistry C Article

Figure 1. (a) Chemical structures of N-polymer and C-polymer using ball-and-stick models. Carbon, nitrogen, and hydrogen atoms are depicted by
green, blue, and gray spheres, respectively. (b) Electrostatic potentials of N-polymer and C-polymer. Experimental CO2 adsorption isotherms per
specific surface area (c) and simulated CO2 adsorption amounts (d) on C-polymer (open symbols) and N-polymer (close symbols) at 243 K (red),
273 K (blue), and 303 K (green). (e) Temperature-dependent adsorption amounts of CO2 on C-polymer (●) and N-polymer (■) in simulations
for 600 bulk CO2 molecules (black) and experiments at 0.1 MPa (green).

functionalities had a weak effect on CO2 capture.44 Kumar et N2 adsorption isotherms were measured at 77 K using a
al. reported that nitrogen-doped microporous carbons had a volumetric apparatus (Autosorb-1, Quantachrome Instru-
marginal effect on CO2 capture in molecular simulations as the ments, Boynton Beach, FL) after vacuum evacuation at 423
micropores played a vital role in capturing CO2.45 A K for more than 2 h. The BET equation was used to calculate
comparable study that uses both experiments and simulations the specific surface area.48 CO2 adsorptions at 243, 273, and
is thus necessary to resolve these inconsistencies. 303 K were also performed.
We performed a comprehensive study of CO2 capture and In addition, 1 mg of both polymers was dispersed in 2 mL of
sensing performances at 243, 273, and 303 K for a nitrogen- isopropyl alcohol using ultrasonication for 1 h. Small amounts
containing organic polymer (named N-polymer) synthesized of these solutions were dropped onto a 10 μm gap indium tin
on carbon nanotubes in experiments and molecular dynamics oxide electrode. The electrodes with samples were heated at
(MD) simulations to evaluate the effect of N atoms on CO2 363 K for 5 min in an N2 atmosphere before measurement of
adsorption. The details of the synthesis of the organic polymer CO2 gas sensing. The four-probe electrical resistivities of the
are reported elsewhere.46 A polymer composed of only carbon samples were measured at 297 K and 5 V. The samples were
atoms (named C-polymer) was also prepared for comparison. exposed to 104 ppm of CO2 gas for 60 s and then refreshed by
introducing N2 gas.
2. EXPERIMENTAL AND SIMULATION PROCEDURES MD simulations were performed using the Large-scale
To synthesize N-polymer (the polymer with N atom doping), Atomic/Molecular Massively Parallel Simulator (LAMMPS)
28 mg of divinyl-1,10-phenathroline (DVP) dissolved in molecular dynamics code. Packmol and moletemplate were
tetrahydrofuran (THF) was impregnated on 10 mg of a used for building topological files.49,50 All-atom N-polymer and
single-walled carbon nanotube (SWCNT) support.47 The C-polymer were composed of 142 and 152 atoms, respectively
impregnated samples were dried at 333 K under reduced (Figure 1a). Five monomer units of C-polymer and four
pressure for 1 h and then sealed in ampule tubes under monomer units of N-polymer were attached by the vinyl group
reduced pressure. The ampule tubes were inserted in a furnace using a head-to-tail connection pattern. We used the
that was preliminarily heated at either 673 or 773 K to prevent semiempirical AM1-BBC method to calculate the charge
polymerization of DVP before vaporization and then kept at distribution in the polymer structures; this method provided
those temperatures for 1 h. The samples were moved to new much more specific distributions of the partial charge in all the
glass tubes and further heated at 413 K for 1 h to remove polymers. During this simulation, the interactions between
possibly remaining precursors. Likewise, to synthesize C- particles were modeled using the standard 12−6 Lennard-
polymer (the polymer without N atom doping), 127.2 mg of Jones (LJ) potential and Coulombic pairwise interaction with a
divinyl-1,10-phenanthrene (DVPA) dissolved in THF was cutoff length of 10.0 Å. For graphene, a 12−6 LJ potential
impregnated on 45.8 mg of a SWCNT support using an model with zero charge was employed by assuming a
incipient wetness impregnation method. The remainder of the honeycomb lattice. The potential parameters were σG = 3.40
preparation method was the same as that used for N-polymer. Å and εG = 0.556 kcal/mol with a C−C distance of 1.42 Å.51
B DOI: 10.1021/acs.jpcc.8b07420
J. Phys. Chem. C XXXX, XXX, XXX−XXX
The Journal of Physical Chemistry C Article

Figure 2. (a) Side-view and top-view snapshots after adsorption of CO2 on N-polymer and C-polymer. (b) Radial distribution functions for CO2
nitrogen (blue), ring (green), and vinyl sites (purple) at 243, 273, and 303 K. The solid curves represent N-polymer, and the dotted curves
represent C-polymer.

The CO2 model was constructed using the TraPPE potential.52 using a semiempirical calculation (AM1) with bond charge
In this model, the CO2 molecule was treated as a linear correction, which was parametrized to reproduce the HF/6-
triatomic molecule with charges placed at the center of each 31G* RESP charge. N-polymer had heterogeneous electro-
atom. The LJ parameters for the C and O atoms from the static charges, whereas C-polymer was neutral. In addition, N-
TraPPE force field were given by σC = 2.8 Å and εC = 0.0537 polymer had rather negative partial charges on N atomic sites
kcal/mol and σO = 3.05 Å and εO = 0.157 kcal/mol, and positive partial charges on the convex edges of
respectively. Moreover, the quadrupole moment of the CO2 phenanthroline rings. The concave edges of the phenanthroline
molecules was considered on the basis of a point charge of and vinyl carbons had small negative charges and neutral
+0.70e placed at the center of mass of the carbon atoms and a charges, respectively. Thus, N-polymer had stronger electro-
charge of −0.35e placed on each oxygen atom. The GAFF static interaction than C-polymer with adsorbed molecules. N2
force field, which is a complete with all available parameters adsorption isotherms at 77 K were measured to determine the
force field with great consistency with the AMBER force field, specific surface areas (see Figure S1a). The shapes of both
was used to construct the all-atom polymer potentials of N- adsorption isotherms in Figure S1a were nearly type II but
polymer and C-polymer.53,54 To confirm the all-mixing LJ showed small adsorption hysteresis, suggesting type IVa
interaction between different types of adsorbates/adsorbents, behavior according to the IUPAC classification.56 These
the Lorentz−Berthelot rules were adopted. results indicate that both polymers had almost nonporous
Two-layer graphene was considered to avoid the effect of surfaces with a minor presence of mesopores. The N2
one side on the other during the adsorption. The two layers of adsorption isotherm on carbon nanotubes was type IVb
(8 nm × 8 nm) graphene with a distance of 3.35 Å were because of the presence of cylindrical nanopores. The specific
composed of 5016 atoms and positioned in the center of a 9 × surface areas of N-polymer and C-polymer obtained from
9 × 22 nm3 unit cell. For C-polymer and N-polymer, 20 and 25 Brunauer−Emmett−Teller (BET) analysis were 30 and 25 m2
molecules were adsorbed on each side of the two-layer g−1, respectively. The surface layer numbers of the polymers
graphene, respectively, in the MD simulation of polymer were evaluated from the decrease of the specific surface areas
adsorption on graphene for 1 ns at 150 K. All the graphene was of the carbon nanotubes. The layer numbers of N-polymer and
considered to be rigid to avoid the slight vibration and C-polymer were 25 and 40, respectively, assuming that the
delocalization of graphene atoms. In total, 200−3000 CO2 molecular areas in the unit molecules were 0.29 and 0.40 nm2,
molecules were randomly positioned in the unit cell with three-
and the polymer type only affected the sample weight.
dimensional periodic boundary conditions. The Nosé−Hoover
However, the layer numbers of the polymers were 2−3 when
algorithm was adopted during the simulation at different
calculated from the weights of the precursors and carbon
constant temperatures of 243, 273, and 303 K. We conducted
nanotubes. This discrepancy indicates that the polymers
2 ns simulations for all the systems to achieve equilibrium. We
induced bundle formation of carbon nanotubes during
considered CO2 molecules within 2 nm of the graphene
surfaces to be adsorbed and the others to be in the bulk. synthesis, leading to a considerable decrease in the surface
areas; thus, the inherent nanopores of the carbon nanotubes
were not beneficial for N2 adsorption by covering with
3. RESULTS AND DISCUSSION polymers. The CO2 adsorption performances were thus
Figure 1a displays the ideal chemical structures of N-polymer discussed using the adsorption amount per surface area.
and C-polymer. N-polymer and C-polymer have zigzag and The CO2 adsorption isotherms were Freundlich-type
linear structures, respectively, and both unit structures are isotherms, as observed in Figure S1b. The CO2 adsorption
phenanthrene forms.55 Figure 1b shows the electrostatic amounts on N-polymer were significantly higher than those on
potentials of both polymers after optimal structure evaluation C-polymer despite the similar specific surface areas. As the
C DOI: 10.1021/acs.jpcc.8b07420
J. Phys. Chem. C XXXX, XXX, XXX−XXX
The Journal of Physical Chemistry C Article

original adsorption isotherms on N-polymer and C-polymer interaction. The carbon ring had a stronger attractive force
hardly compared with each other, the CO2 adsorption than the vinyl group, estimated from the higher atomic density.
isotherms per specific surface area were calculated using the In addition, the CO2−vinyl peaks for C-polymer disappeared
adsorption isotherms and specific surface area, as shown in because of the weak attractive force of the vinyl group. The
Figure 1c. At 0.1 MPa, N-polymer adsorbed 1.5, 1.2, and 0.9 radial distribution peaks between CO2 adsorption sites
mg m−2 at 243, 273, and 303 K, respectively. The decreased with increasing temperature, whereas these peaks
corresponding adsorbed amounts on C-polymer were 1.1, hardly changed for N-polymer.
0.7, and 0.3 mg m−2, respectively. The CO2 adsorption The temperature dependence of the dynamic properties
amounts increased with decreasing temperature for both N- were evaluated from the mean-square displacement and self-
polymer and C-polymer, which is a well-known phenomenon diffusion coefficient, although temperature dependence of the
for most adsorbents. N-polymer apparently exhibited better static structure was rarely observed. The mean-square
CO2 adsorption performance than C-polymer, with increases displacement in Figure S4 was calculated from the CO2
of 40%, 70%, and 200% at 243, 273, and 303 K, respectively. trajectory in MD simulations of 3000 CO2 molecules in the
This finding indicates that strong adsorption sites were formed systems. CO2 in the vicinity of nitrogen adsorption sites of N-
for N-polymer. The simulated CO2 adsorptions were evaluated polymer had smaller mean-square displacements than those
from the MD simulation of CO2 on a polymer fabricated on surrounding ring and vinyl sites of N-polymer and C-polymer.
graphene (Figure 1d). The simulated adsorption curves The mean-square displacements surrounding ring and vinyl
exhibited the same trend of temperature dependence and sites of N-polymer and C-polymer, as shown in Figure S4,
enhancement of CO2 adsorption on N-polymer. Figure 1e increased with increasing temperature, whereas that surround-
shows the temperature dependences of CO2 adsorption for the ing nitrogen sites was relatively constant. CO2 adsorbed on the
experiments and simulations. The experimental and simulated nitrogen sites of N-polymer was strongly connected to nitrogen
amounts were obtained at 0.1 MPa and for 600 molecules in atoms in the temperature range, whereas CO2 adsorbed on
the bulk, respectively. The adsorption amounts clearly ring and vinyl sites was released above 273 K and the mean-
decreased with increasing temperature for both N-polymer square displacement was close to that of C-polymer. The self-
and C-polymer. However, the rates of decrease were moderate diffusion coefficients calculated from the following equation
in N-polymer, indicating the strong adsorption potential of N- clearly describe the CO2 dynamics, as shown in Figure 3.
polymer. The difference between the CO2 adsorptions on N-
polymer and C-polymer at higher temperature was thus larger D = ⟨(R(t ) − R(0))2 ⟩/6t (1)
than that at lower temperature. The CO2 adsorptions on N-
polymer were most likely retained because of the Coulomb
interaction between CO2 and N-polymer.
Figure 2a presents snapshots of the CO2 adsorbed on N-
polymer and C-polymer (see also the snapshots in Figures S2
and S3). CO2 was uniformly distributed on both polymers in
the side views. However, the nitrogen and ring sites of N-
polymer dominantly adsorbed CO2 in the top views, whereas
CO2 was mainly adsorbed at the ring sites of C-polymer. Here,
N-polymer had three adsorption sites on nitrogen, ring, and
vinyl groups, whereas C-polymer only had ring and vinyl
adsorption sites. The radial distribution functions between
CO2 and the adsorption sites were calculated to determine the
active site for CO2 adsorption, as shown in Figure 2b. The
nearest-neighbor distances for CO2−nitrogen, −ring, and− Figure 3. Self-diffusion coefficients of CO2 near ring and vinyl sites of
C-polymer (blue) and near nitrogen sites (red) and ring and vinyl
vinyl groups of N-polymer were 0.31, 0.35, and 0.37 nm, sites (green) of N-polymer at 243, 273, and 303 K.
respectively. However, the nearest-neighbor distances for
CO2−ring and −vinyl groups of C-polymer were 0.34 and
0.37 nm, respectively. The distance between CO2 and a where D is the self-diffusion coefficient and R(t) is the position
nitrogen group was the shortest because the nitrogen site had vector of CO2 at time t. The self-diffusion coefficients of CO2
the largest negative partial charge, strongly attracting the in the vicinity of nitrogen sites of N-polymer moderately
positive charge of CO2 and a CO2 molecule was preferentially increased with increasing temperature unlike those for the ring
positioned in the middle of two nitrogen atoms from the and vinyl sites. The self-diffusion coefficients near the ring and
snapshots. The distances between CO2 and the ring and vinyl vinyl sites of C-polymer were higher than those of N-polymer.
groups were similar for N-polymer and C-polymer; however, However, the self-diffusion coefficients in the vicinity of ring
the CO2−ring distance for C-polymer was slightly shorter than and vinyl sites of N-polymer and C-polymer were indistin-
that for N-polymer because the ring periphery space of C- guishable at 303 K. The kinetic energy of CO2 overcame the
polymer was larger than that of N-polymer after CO2 weak interaction between CO2 and ring or vinyl sites mainly by
adsorption on nitrogen sites. Carbon atoms on ring groups Coulomb interaction. However, the nitrogen sites strongly
of C-polymer had slight negative charges, whereas carbon attracted CO2, which was associated with a reduction of kinetic
atoms for N-polymer had nonuniform charges, as shown in energy. The negatively charged nitrogen sites of N-polymer
Figure 1b. Despite the different charge distributions, the thus played a vital role in decreasing the surface self-diffusion
distance between CO2 and carbon groups was unchanged, coefficient of CO2 and maintaining the CO2 adsorption
suggesting that CO2 adsorption on carbon groups was amounts. The restriction of CO2 dynamics on nitrogen sites
dominated by a dispersion interaction rather than a Coulomb led to considerable CO2 adsorption even at higher temper-
D DOI: 10.1021/acs.jpcc.8b07420
J. Phys. Chem. C XXXX, XXX, XXX−XXX
The Journal of Physical Chemistry C Article

ature, whereas carbon sites such as ring and vinyl groups were polymer during CO2 adsorption and desorption and from self-
available for inducing CO2 adsorption at lower temperature, heating due to the low resistivity during the measurements.
which is consistent with the CO2 adsorption isotherms in Structural change of the flexible coordination polymer was
Figure 1c−e. induced by CO2 adsorption and desorption, and its adsorption
Significant CO2 adsorption and restriction of CO2 dynamics changes gradually decreased with repeating cycles.59 The
were observed on N-polymer, and CO2 was reversibly structural changes of N-polymer and C-polymer were thus
adsorbed. Strong and reversible adsorption of CO2 are useful evaluated from the mean-square displacements of the
for CO2 sensors. Figure S5a shows the resistance changes with polymers, as shown in Figure S6. The structure change of C-
time in CO2 and N2 atmospheres. The resistances of both N- polymer was much greater than that of N-polymer during CO2
polymer and C-polymer increased with CO2 introduction. CO2 adsorption. We thus concluded that the slight decrease of the
can be a weak Lewis acid with acceptance of an electron pair CO2 sensor performance on C-polymer was caused by its
from a nitrogen group of N-polymer or from the ring and vinyl flexibility rather than self-heating, as the difference between the
groups of C-polymer. Both polymers were thus n-type resistances of N-polymer and C-polymer was relatively small.
semiconductors. The relative resistance changes were calcu-
lated from the differential between the resistances at t and 0 s, 4. CONCLUSION
as shown in Figure 4. The relative resistance change of C-
We demonstrated the correlation between CO2 and nitrogen
atoms on a polymer by evaluating the CO2 adsorption
properties on a nitrogen-doped polymer and a simple polymer
as well as their sensing performance. N-polymer possessed a
considerable partial charge distribution, including a negative
charge on a nitrogen site, whereas the partial charges of C-
polymer were almost neutral. Although a CO2 molecule has
both positive and negative partial charges, CO2 is considered
to work as a weak Lewis acid and to be strongly attracted to
nitrogen sites. CO2 adsorption isotherms at 243, 273, and 303
K on N-polymer revealed higher CO2 adsorption uptakes than
those for C-polymer; these findings were also suggested by
MD simulations. CO2 was mainly adsorbed on the nitrogen
sites of N-polymer at 303 K but fairly homogeneously
distributed at 243 and 273 K. The self-diffusion coefficient
of CO2 on the nitrogen sites in the MD simulation was
Figure 4. Response curves to CO2 introduction on N-polymer and C- apparently smaller than those on ring and vinyl sites, and the
polymer. self-diffusion coefficient remained mostly constant at high
temperature. The CO2 gas sensor performances indicated the
higher sensitivity of CO2 on N-polymer because of the strong
polymer was ∼2%, which is similar to that for reported carbon attraction force with nitrogen sites and adequate reversibility
sensors.57,58 A resistance change of up to 10% was achieved for because of the reduced structural change of N-polymer.
N-polymer, which was 5 times larger than that for C-polymer. Therefore, nitrogen doping on the polymer improved the
The large resistance changes were caused by CO2 adsorption reversibility and capacity of CO2 adsorption as well as the
on nitrogen sites inducing electron transfer from a nitrogen site electrical conductivity, yielding high CO2 sensor performance.


to CO2. The first steps of the resistance changes of
approximately 4% and 0.5% with CO2 introduction were
observed for N-polymer and C-polymer, respectively. The first
ASSOCIATED CONTENT
step of the resistance change was observed instantaneously *
S Supporting Information
after switching gases. The electrical resistance could be The Supporting Information is available free of charge on the
changed by changing the pressure as well as through ACS Publications website at DOI: 10.1021/acs.jpcc.8b07420.
adsorption. We expected rapid and slow responses for the
pressure change and adsorption, respectively. Thus, the first N2 and CO2 adsorption isotherms on polymers,
and second steps were mainly attributed to pressure fluctuation snapshots of CO2 on N-polymer and C-polymer,
and adsorption, respectively. Figure S5b shows the normalized mean-square displacement of CO2, electrical resistivity
relative resistance by the top of the first peak. The recovering changes of polymers by CO2 insertion, and mean-square
curves of the resistance by N2 introduction for N-polymer were displacement of polymers during CO2 adsorption (PDF)


slower than those for C-polymer because of the restriction of
CO2 dynamics on nitrogen sites, as expected from the MD AUTHOR INFORMATION
simulation results in Figure 3. However, the normalized
resistance peaks were maintained for N-polymer, whereas they Corresponding Author
were gradually reduced for C-polymer. This finding was *E-mail: ohba@chiba-u.jp (T.O.).
unexpected as chemisorption and/or strong physical adsorp- ORCID
tion typically impose a reduction of resistance change. CO2
Yasuhiro Yamada: 0000-0002-4974-5148
was thus strongly adsorbed on N-polymer, although reversible
adsorption was observed. The excellent cyclability might have Tomonori Ohba: 0000-0001-8207-3630
originated from maintenance of the adsorption amounts in the Notes
CO2 environment by avoidance of structural change of N- The authors declare no competing financial interest.
E DOI: 10.1021/acs.jpcc.8b07420
J. Phys. Chem. C XXXX, XXX, XXX−XXX
The Journal of Physical Chemistry C Article

■ ACKNOWLEDGMENTS
The authors thank the Supercomputer Center, Institute for
(17) Wang, Y.; LeVan, M. D. Adsorption Equilibrium of Carbon
Dioxide and Water Vapor on Zeolites 5A and 13X and Silica Gel:
Pure Components. J. Chem. Eng. Data 2009, 54, 2839−2844.
Solid State Physics, University of Tokyo for use of the facilities. (18) Choi, S.; Drese, J. H.; Jones, C. W. Adsorption Materials for
This research was supported by the Japan Society for the Carbon Dioxide Capture from Large Anthropogenic Point Sources.
Promotion of Science KAKENHI (Grants 26706001 and ChemSusChem 2009, 2, 796−854.
15K12261), ESPEC Foundation for Global Environment (19) Akhtar, F.; Liu, Q.; Hedin, N.; Bergstrom, L. Strong and Binder
Research and Technology, and the New Energy and Industrial Free Structure Zeolite Sorbents with Very High CO2-over-N2
Technology Development Organization (NEDO) under the Selectivities and High Capacities to Adsorb CO2 Rapidlly. Energy
Environ. Sci. 2012, 5, 7664−7673.
Ministry of Economy Trade and Industry of Japan (Grants (20) Foo, M. L.; Matsuda, R.; Hijikata, Y.; Krishna, R.; Sato, H.;
16101400-0 and 16101401-0). Horike, S.; Hori, A.; Duan, J.; Sato, Y.; Kubota, Y.; Takata, M.;

■ REFERENCES
(1) Nandi, M.; Uyama, H. Exceptional CO2 Adsorbing Materials
Kitagawa, S. An Adsorbate Discriminatory Gate Effect in a Flaxible
Porous Coordination Polymer for Selective Adsorption of CO2 over
C2H2. J. Am. Chem. Soc. 2016, 138, 3022−3030.
(21) Comotti, A.; Fraccarollo, A.; Bracco, S.; Beretta, M.; Distefano,
Under Different Conditions. Chem. Rec. 2014, 14, 1134−1148.
G.; Cossi, M.; Marchese, L.; Riccardi, C.; Sozzani, P. Porous
(2) Haszeldine, R. S. Carbon Capture and Storage: How Green Can
Dipeptide Crystals as Selective CO2 Adsorbents: Experimental
Black Be? Science 2009, 325, 1647−1652.
Isotherm vs. Grand Canonical Mote Carlo Simulations and MAS
(3) Monastersky, R. Global Carbon Dioxide Levels Near Worrisome
NMR Spectroscopy. CrystEngComm 2013, 15, 1503−1507.
Milestone. Nature 2013, 497, 13−14. (22) Hicks, J. C.; Drese, J. H.; Fauth, D. J.; Gray, M. L.; Qi, G.;
(4) Hinkov, I.; Lamari, F. D.; Langlois, P.; Dicko, M.; Chilev, C.;
Jones, C. W. Designing Adsorbents for CO2 Capture from Flue Gas-
Pentchev, I. Carbon Dioxide Capture by Adsorption. J. Chem. Technol. Hyperbranched Aminosilicas Capable of Capturing CO2 Reversibly. J.
Metall. 2016, 51, 609−626. Am. Chem. Soc. 2008, 130, 2902−2903.
(5) White, C. M.; Smith, D. H.; Jones, K. L.; Goodman, A. L.; Jikich, (23) Khatri, R. A.; Chuang, S. S. C.; Soong, Y.; Gray, M. Thermal
S. A.; LaCount, R. B.; Ozdemir, E.; Morsi, B. I.; Schroeder, K. T. and Chemical Stability of Regenerable Solid Amine Sorbent for CO2
Sequestration of Carbon Dioxide in Coal with Enhanced Coalbed Capture. Energy Fuels 2006, 20, 1514−1520.
Methane Recovery, A Review. Energy Fuels 2005, 19, 659−724. (24) Watanabe, T.; Khan, S. M.; Kanoh, H.; Ohba, T. Significant
(6) Wang, W.; zhou, M.; Yuan, D. Carbon Dioxide Capture in CO2 Adsorption Ability of Nanoscale BaTiO3 Ceramics Fabricated by
Amorphous Porous Organic Polymers. J. Mater. Chem. A 2017, 5, Carbon-Template-Solvothermal Reactions. Phys. Chem. Ind. J. 2017,
1334−1347. 12, 101.
(7) Liu, L.; Bhatia, S. K. Molecular Simulation of CO2 Adsorption in (25) An, H.; Feng, B.; Su, S. CO2 Capture by Electrothermal Swing
the Presence of Water in Single-Walled Carbon Nanotubes. J. Phys. Adsorption with Activated Carbon Fibre Materials. Int. J. Greenhouse
Chem. C 2013, 117, 13479−13491. Gas Control 2011, 5, 16−25.
(8) Himeno, S.; Komatsu, T.; Fujita, S. High-Pressure Adsorption (26) Hutson, N. D.; Speakman, S. A.; Payzant, E. A. Structural
Equilibrium of Methene and Carbon Dioxide on Several Activated Effects on the High Temperature Adsorption of CO2 on a Synthetic
Carbons. J. Chem. Eng. Data 2005, 50, 369−376. Hydrotalcite. Chem. Mater. 2004, 16, 4135−4143.
(9) Hao, G. P.; Li, W. C.; Lu, A. H. Novel Porous Solids for Carbon (27) Takamura, Y.; Narita, S.; Aoki, J.; Hironaka, S.; Uchida, S.
Dioxide Capture. J. Mater. Chem. 2011, 21, 6447−6451. Evaluation of Dual-Bed Pressure Swing Adsorption for CO2 Recovery
(10) Zhu, X.; Hillesheim, P. C.; Mahurin, S. M.; Wang, C.; Tian, C.; from Boiler Exhust Gas. Sep. Purif. Technol. 2001, 24, 519−528.
Brown, S.; Luo, H.; Veith, G. M.; Han, K. S.; Hagaman, E. W.; Liu, (28) Grande, C. A.; Rodrigues, A. E. Electric Swing Adsorption for
H.; Dai, S. Efficient CO2 Capture by Porous, Nitrogen Doped CO2 Removal from Flue Gases. Int. J. Greenhouse Gas Control 2007, 2,
Carbonaceous Adsorbents Derived from Task-specific Ionic Liquids. 194−202.
ChemSusChem 2012, 5, 1912−1917. (29) Ben, T.; Ren, H.; Ma, S.; Cao, D.; Lan, J.; Jing, X.; Wang, W.;
(11) Zhao, L.; Bacsik, Z.; Hedin, N.; Wei, W.; Sun, Y.; Antonietti, Xu, J.; Deng, F.; Simmons, J. M.; Qiu, S.; Zhu, G. Targeted Synthesis
M.; Titirici, M. M. Carbon Dioxide Capture on Amine-Rich of a Porous Aromatic Framework with High Stability and Exceptional
Carbonaceous Materials Derived from Glucose. ChemSusChem High Surface Area. Angew. Chem., Int. Ed. 2009, 48, 9457−9460.
2010, 3, 840−845. (30) Ren, H.; Ben, T.; Sun, F.; Guo, M.; Jing, X.; Ma, H.; Cai, K.;
(12) Shao, X.; Feng, Z.; Xue, R.; Ma, C.; Wang, W.; Peng, X.; Cao, Qiu, S.; Zhu, G. Synthesis of a Porous Aromatic Framework for
D. Adsorption of CO2, CH4, CO2/N2 and CO2/CH4 in Novel Adsorbing Organic Pollutants Application. J. Mater. Chem. 2011, 21,
Activated Carbon Based: Preparation, Measurements and Simulation. 10348−10353.
AIChE J. 2011, 57, 3042−3051. (31) Ben, T.; Pei, C.; Zhang, D.; Xu, J.; Deng, F.; Jing, X.; Qiu, S.
(13) Wu, Z.; Hao, N.; Xiao, G.; Liu, L.; Webley, P.; Zhao, D. One- Gas Storage in Porous Aromatic Frameworks (PAFs). Energy Environ.
pot Generation of Mesoporous Carbon Supported Nanocrystalline Sci. 2011, 4, 3991−3999.
Calcium Oxides Capable of Efficient CO2 Capture Over a Wide (32) Ren, H.; Ben, T.; Wang, E.; Jing, X.; Xue, M.; Liu, B.; Cui, Y.;
Range of Temperatures. Phys. Chem. Chem. Phys. 2011, 13, 2495− Qiu, S.; Zhu, G. Targated Synthesis of a 3D Porous Arometic
2503. Framework for Selective Sorption of Benzene. Chem. Commun. 2010,
(14) Sumida, K.; Rogow, D. L.; Mason, J. A.; McDonald, T. M.; 46, 291−293.
Bloch, E. D.; Herm, Z. R.; Bae, T. H.; Long, J. R. Carbon Dioxide (33) Yuan, Y.; Sun, F.; Ren, H.; Jing, X.; wang, W.; Ma, H.; Zhao,
Capture in Metal-Organic Frameworks. Chem. Rev. 2012, 112, 724− H.; Zhu, G. Targeted Synthesis of a Porous Aromatic Framework with
781. a High Adsorption Capacity for Organic Molecules. J. Mater. Chem.
(15) Fracaroli, A. M.; Furukawa, H.; Suzuki, M.; Dodd, M.; Okajima, 2011, 21, 13498−13502.
S.; Gandara, F.; Reimer, J. A.; Yaghi, O. M. Metal-Organic (34) Cong, H. L.; Zhang, M.; Chen, Y.; Chen, K.; Hao, Y.; Zhao, Y.;
Frameworks with Precisely Designed Interior for Carbon Dioxide Feng, L. Highly Selective CO2 Capture by Nitrogen Enriched Prous
Capture in the Presence of Water. J. Am. Chem. Soc. 2014, 136, 8863− Carbon. Carbon 2015, 92, 297−304.
8866. (35) Srinivas, G.; Krungleviciute, V.; Guo, Z.; Yildirim, T.
(16) Cavenati, S.; Grande, C. A.; Rodrigues, A. E. Adsorption Exceptional CO2 Capture in a Hierarchically Porous Carbon with
Equilibrium of Methene, Carbon Dioxide, and Nitrogen on Zeolite Simultaneous High Surface Area and Pore Volume. Energy Environ.
13X at High Pressure. J. Chem. Eng. Data 2004, 49, 1095−1101. Sci. 2014, 7, 335−342.

F DOI: 10.1021/acs.jpcc.8b07420
J. Phys. Chem. C XXXX, XXX, XXX−XXX
The Journal of Physical Chemistry C Article

(36) Wang, J.; Senkovska, I.; Oschatz, M.; Lohe, M. R.; Borchardt, (55) Wang, L.; Yang, R. T. Significantly Increased CO2 Adsorption
L.; Heerwig, A.; Liu, Q.; Kaskel, S. Imine-Linked Polymer-Derived Performance of Nanostructured Templated Carbon by Tuning
Nitrogen-Doped Microporous Carbons with Excellent CO2 Capture Surface Area and Nitrogen Doping. J. Phys. Chem. C 2012, 116,
Properties. ACS Appl. Mater. Interfaces 2013, 5, 3160−3167. 1099−1106.
(37) Chen, Q.; Luo, M.; Hammershoj, P.; Zhou, D.; Han, Y.; (56) Thommes, M.; Kaneko, K.; Neimark, L. V.; Olivier, J. P.;
Laursen, B. W.; Yan, C. G.; Han, B. H. Microporous Polycarbazole Rodriguez-Reinoso, F.; Rouquerol, J.; Sing, K. S. W. Physisorption of
with High Specific Surface Area for Gas Storage and Separation. J. Gases, with Special Reference to the evaluation of Surface Area and
Am. Chem. Soc. 2012, 134, 6084−6087. Pore Size Distribution (IUPAC Technical Report). Pure Appl. Chem.
(38) Hug, S.; Mesch, M. B.; Oh, H.; Popp, N.; Hirscher, M.; Senker, 2015, 87, 1051−1069.
J.; Lotsch, B. V. A Fluorene Based Covalent Triazine Franework with (57) Zhou, Y.; Jiang, Y. D.; Xie, T.; Tai, H. L.; Xie, G. Z. A Novel
High CO2 and H2 Capture and Storage Capacities. J. Mater. Chem. A Sensing Mechanism for Resistive Gas Sensors Based on Layered
2014, 2, 5928−5936. Reduced Graphene Oxide Thin Films at Room Temperature. Sens.
(39) Rabbani, M. G.; El-Kaderi, H. M. Synthesis and Character- Actuators, B 2014, 203, 135−142.
ization of Porous Benzimidazole Linked Polymers and Their (58) Fan, X.; Elgammal, K.; Smith, A. D.; Ö stling, M.; Delin, A.;
Performance in Small Gas Storage and Selective Uptake. Chem. Lemme, M. C.; Niklaus, F. Humidity and CO2 Gas Sensing Properties
Mater. 2012, 24, 1511−1517. of Double-Layer Graphene. Carbon 2018, 127, 576−587.
(40) Rabbani, M. G.; El-Kaderi, H. M. Template-Free Synthesis of a (59) Choi, H. S.; Suh, M. P. Highly Selective CO2 Capture in
Highly Porous Benzimidazole-Linked Polymer for CO2 Capture and Flexible 3D Coordination Polymer Networks. Angew. Chem., Int. Ed.
2009, 48, 6865−6869.
H2 Storage. Chem. Mater. 2011, 23, 1650−1653.
(41) Li, Q.; Yang, J.; Feng, D.; Wu, Z.; Wu, Q.; Park, S. S.; Ha, C.;
Zhao, D. Facile Synthesis of Porous Carbon Nitride Spheres with
Hierarchical Three-Dimensional Mesostructures for CO2 Capture.
Nano Res. 2010, 3, 632−642.
(42) Maroto-Valer, M. M.; Tang, Z.; Zhang, Y. CO2 Capture by
Activated and Impregnated Anthracites. Fuel Process. Technol. 2005,
86, 1487−1502.
(43) Xing, W.; Liu, C.; Zhou, Z.; Zhang, L.; Zhou, J.; Zhuo, S.; Yan,
Z.; Gao, H.; Wang, G.; Qiao, S. Z. Superior CO2 Uptake of N-doped
Activated Carbon Through Hydrogen-bonding Interaction. Energy
Environ. Sci. 2012, 5, 7323−7327.
(44) Sevilla, M.; Parra, J. B.; Fuertes, A. B. Assesment of the Role of
Micropore Size and N-Doping in CO2 Capture by Porous Carbons.
ACS Appl. Mater. Interfaces 2013, 5, 6360−6368.
(45) Kumar, K. V.; Preuss, K.; Lu, L.; Guo, Z. X.; Titirici, M. M.
Effect of Nitrogen Doping on the CO2 Adsorption Behavior in
Nanoporous Carbon Structures: A Molecular Simulation Study. J.
Phys. Chem. C 2015, 119, 22310−22321.
(46) Yamada, Y.; Gohda, S.; Abe, K.; Togo, T.; Shimano, N.; Sasaki,
T.; Tanaka, H.; Ono, H.; Ohba, T.; Kubo, S.; Ohkubo, T.; Sato, S.
Carbon Materials with Controlled Edge Structures. Carbon 2017, 122,
694−701.
(47) Hata, K.; Futaba, D. N.; Mizuno, K.; Namai, T.; Yumura, M.;
Iijima, S. Water-Assisted Highly Efficient Synthesis of Impurity-Free
Single-Walled Carbon Nanotubes. Science 2004, 306, 1362−1364.
(48) Barrett, E. P.; Joyner, L. G.; Halenda, P. P. The Determination
of Pore Volume and Area Distributions in Porous Substances. I.
Computaions from Nitrogen Isotherms. J. Am. Chem. Soc. 1951, 73,
373−380.
(49) Martinez, L.; Andrade, R.; Birgin, E. G.; Martinez, J. M.
PackMOL: A Package for Building Initial Configurations for
Molecular Dynamics Simulations. J. Comput. Chem. 2009, 30,
2157−2164.
(50) Martinez, J. M.; Martinez, L. Packing Optimization for
Automated Generation of Complex System̀ s Initial Configurations
for Molecular Dynamics and Docking. J. Comput. Chem. 2003, 24,
819−825.
(51) Ohba, T. The Thinnest Molecular Seperation Sheet by
Graphene Gates of Single-Walled Carbon Nanohorns. ACS Nano
2014, 8, 11313−11319.
(52) Potoff, J. J.; Siepmann, J. I. Vapor-Liquid Equilibria of Mixtures
Containing Alkanes, Carbon Dioxide and Nitrogen. AIChE J. 2001,
47, 1676−1682.
(53) Wang, J.; Wang, W.; Kollman, P. A.; Case, D. A. Automatic
Atom Type and Bond Type Perception in Molecular Mechanical
Calculations. J. Mol. Graphics Modell. 2006, 25, 247−260.
(54) Wang, J.; Wolf, R. M.; Caldwell, J. W.; Kollman, P. A.; Case, D.
A. Development and Testing of General Amber Force Field. J.
Comput. Chem. 2004, 25, 1157−1174.

G DOI: 10.1021/acs.jpcc.8b07420
J. Phys. Chem. C XXXX, XXX, XXX−XXX

You might also like