Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

PCCP

View Article Online


PAPER View Journal | View Issue

Energy storage properties of a two-dimensional


TiB4 monolayer
Published on 25 May 2019. Downloaded by University of Frankfurt on 7/26/2019 4:59:43 AM.

Cite this: Phys. Chem. Chem. Phys.,


2019, 21, 13151 ab
Zhiyang Liu, Erdong Wu,a Jiangxu Liab and Shi Liu *a

Herein, the energy storage properties of TiB4 monolayers were studied within the density functional
theory framework. Both CH4 and H2 were chosen as adsorption molecules, and their interactions with a
TiB4 sheet were investigated. TiB4 attracted gas molecules via open Ti sites, and each Ti atom could
adsorb a maximum of two molecules. Via the electronic density of the states and atomic charge
analysis, we found that the mechanism for gas adsorption was mainly electrostatic. For H2 adsorption
cases, orbital interactions also made contributions. As the combustion energy of one CH4 molecule is
three times that of one H2, the TiB4–2CH4 compound can achieve the best equivalent gravimetric
Received 3rd April 2019, hydrogen density of 10.14 wt% with the average adsorption energy of 0.38 eV. Ab initio molecular
Accepted 24th May 2019 dynamics calculations on this compound showed that there was no kinetic barrier during CH4
DOI: 10.1039/c9cp01864f desorption. Moreover, the stacking of the TiB4 monolayers could weaken the energy storage capacity.
Therefore, it should be avoided in practial usage. Based on the abovementioned results, the TiB4 mono-
rsc.li/pccp layer was suggested to be a promising candidate for onboard energy storage.

1. Introduction an obvious limitation: re-hydrogenation occurs only under high


temperature and hydrogen pressure conditions.8 The storage of
Energy is a main issue facing humanity today. Compared with hydrogen in molecular form could be an effective way to deal
traditional fossil fuels, hydrogen is a clean, efficient and renew- with the abovementioned problems due to no bond formation
able energy source; thus, it can be an efficient source for the or breaking during the process of usage. Many calculations
next generation of fuels.1 The challenge in hydrogen utilization have been conducted on nanomaterials9–20 and framework
is the storage of hydrogen with high gravimetric and volumetric materials21–26 for molecular hydrogen storage due to their high
density under near ambient thermodynamic conditions. Hydro- surface/volume ratios. For instance, Duan et al.17 investigated
gen can be stored in two forms: atomic and molecular. As a Li-decorated CN via simulation and achieved the theoretical
common and typical atomic hydrogen storage material, gravimetric hydrogen density of 10.81 wt%. Yang et al.14 simu-
magnesium can store up to 7.7 wt% hydrogen; however, the lated Ti-decorated g-C3N4 and achieved the hydrogen capacity
hydrogen desorption temperature of MgH2 is high, and the of 9.70 wt%. However, considering the experimental results of
(de)hydrogenation kinetics is poor; thus, additives are needed these materials obtained to date, most storage capacities
to improve these properties.2,3 Bulk alloys, such as vanadium obtained at ambient temperature and pressure still failed to
alloys,4–6 store hydrogen by forming metal hydrides, and the reach the DOE target.27–30 As a result, molecular hydrogen
stored hydrogen can be used at around room temperature and storage materials need to be further explored.
standard atmospheric pressure. However, they usually undergo Recently, a new kind of 2D material, i.e. TiB4, has been
chemical reactions during storage, and bonds are broken or predicted.31 It has much lower weight than TM-TCNQ,32,33 and
formed during hydrogen usage; moreover, the hydrogen per- a similar open metal site that may attract gas molecules.
cent is too far to reach the target of 5.5 wt% (ultimate 6.5 wt%) Considering the valences of TM, one Sc (3d14s2) atom can
hydrogen by 2025 according to the US Department of Energy coordinate with three B atoms to form a carbon-like electronic
(DOE);7 in this regard, a covalently bonded compound, i.e. shell—ScB3 nanotube, which is expected to be a stable hydro-
LiBH4, can be a promising hydrogen storage material because gen storage material.34 Then, the TiB4 nanostructure contain-
of its high hydrogen storage capacity (18.4 wt%); however, it has ing one Ti (3d24s2) atom and four B atoms could also be stable.
Via thermodynamic calculations, Qu and coworkers found that
a
the TiB4 sheet still remained stable at 2500 K,31 much higher
Institute of Metal Research, Chinese Academy of Sciences, 72 Wenhua Road,
Shenyang, 110016, China. E-mail: sliu@imr.ac.cn
than the thermal stability of isoreticular MOFs (673 K),35 which
b
School of Materials Science and Engineering, University of Science and Technology could be conducive to the actual usage. Therefore, in this study,
of China, 72 Wenhua Road, Shenyang, 110016, China we chose TiB4 as a substrate material. For energy storage, CH4

This journal is © the Owner Societies 2019 Phys. Chem. Chem. Phys., 2019, 21, 13151--13156 | 13151
View Article Online

Paper PCCP

could be another choice due to its low cost and limited air ETiB4ng are the total energy of the unit TiB4 sheet and its counter-
pollution upon combustion. Although the CH4 molecules have parts with (n 1) and n adsorbed gas molecules, respectively.
larger repulsion force than H2, the energy of one CH4 is
3.1 times that of one H2 molecule. Then, the addition of CH4
could also increase the energy capacity of the storage system. 3. Results and discussion
For example, Wu and coworkers found that the mixture of CH4
and H2 significantly enhanced the energy storage capacity of The optimized structure of the TiB4 monolayer is shown in
Li-decorated graphene and Sc-decorated armchair graphene Fig. 1(a). The calculated lattice constants of the unit cell were
nanoribbons.36 Therefore, herein, via first principles calcula- a = b = 4.10 Å with the Ti–B bond distance of 2.22 Å, the B–B
tions, not only the hydrogen storage properties of TiB4 were bond length in the B squares of 1.71 Å, and the B–B bond
Published on 25 May 2019. Downloaded by University of Frankfurt on 7/26/2019 4:59:43 AM.

studied, but also CH4 adsorption on the TiB4 monolayer was length in the sharing wheel edges of 1.68 Å. These values were
considered. in accordance with the previously reported values of 4.12 Å,
2.22 Å, 1.72 Å and 1.68 Å.31 The electron localization function
(ELF)40 was used to analyze the electron distribution and is
2. Computational methods shown in Fig. 1(b). The regions where the ELF values were 0 and
Density functional theory (DFT) calculations were performed 1 corresponded to no and fully localized electrons, respectively.
using the Vienna ab initio simulation package (VASP)37–39 and It can be seen from Fig. 1(b) that a few electrons are distributed
the projector augmented-wave (PAW)40,41 method. The Perdew– around the Ti atoms; this indicates the ionic interaction
Burke–Ernzerhof (PBE)42 functional within the generalized between Ti and the B network. Electrons accumulated in the
gradient approximation (GGA) was used to determine the middle of the B–B bonds; this confirmed the presence of strong
exchange–correlation potential. The DFT-D3 scheme of covalent states between B atoms. Phonon dispersion calculations46
Grimme for the vdW correction43–45 was applied to investigate were performed to evaluate our optimized TiB4 monolayer, and no
the gas adsorption of H2 and CH4 on the TiB4 sheet. The imaginary frequency existed in the first Brillouin zone (Fig. 1(c));
original structure parameters of TiB4 chosen herein were this confirmed the stability of this monolayer.
obtained from the literature.31 A vacuum region of more than For gas adsorption on the TiB4 monolayer, at first, we
20 Å was used to avoid the interaction between adjacent layers. considered the adsorption of a single gas molecule (H2 or
Fully optimized structures were obtained when the energy and CH4) and varied its position above the surface of the TiB4 unit.
maximum force were less than 10 5 eV and 0.01 eV Å 1, Via optimization, we found that the preferred adsorption site
respectively. The kinetic energy cut-off for the electron wave was above the Ti atom, and the gas structures changed after
functions was 500 eV. We used 12  12  1 G-centered adsorption. The optimized equilibrium configurations of the
Monkhorst–Pack k-point samplings in the Brillouin zone for gas-adsorbed TiB4 monolayers are shown in Fig. 2(a)–(f), corres-
geometry optimization and property calculation. ponding to the side and top views. When one H2 adsorbed at
The average gas adsorption energy (Egave) and the continuous the Ti site (Fig. 2(b)), its H–H bond length increased from 0.75 Å
adsorption energy (Egcon) were defined as follows: to 0.79 Å. For the adsorption of one CH4 (Fig. 2(d)), two H atoms
of the CH4 molecule stayed close to the Ti site with two
Egave = (nEg + ETiB4 ETiB4ng)/n (1) corresponding C–H bond lengths of 1.11 Å. Moreover, the other
two H atoms were far away from the TiB4 monolayer, and their
Egcon = Eg + ETiB4(n 1)g ETiB4ng (2)
C–H bond lengths were 1.09 Å. As more gas molecules were
where g represents the gas molecule, Eg is the total energy of one added to the TiB4 surface, we found that at each side of the
H2 or CH4 molecule in the gas phase, and ETiB4, ETiB4(n 1)g, and open Ti site, only one gas molecule stably adsorbed. When two

Fig. 1 (a) The top view of a 2  2  1 TiB4 super cell and (b) the corresponding ELF map sliced perpendicular to the (001) direction. (c) Phonon dispersion
of the TiB4 monolayer. G(0.0 0.0 0.0), X(0.0 0.5 0.0), and M(0.5 0.5 0.0) refer to special q-points in the first Brillouin zone of the reciprocal space. Ti and B
atoms are denoted by blue and green spheres, respectively. The black dashed square marks a unit cell.

13152 | Phys. Chem. Chem. Phys., 2019, 21, 13151--13156 This journal is © the Owner Societies 2019
View Article Online

PCCP Paper

Table 1 The adsorption energy (Egave and Egcon, eV per g), Ti–g distance
(dTi–g, Å) and Bader charges (e) of the mH2–nCH4–TiB4 compounds. g
stands for H2 or CH4

dTi–g
Host Egave Egcon 1st g 2nd g QTi QB QH2 QCH4
TiB4 +1.51 0.38
TiB41H2 0.54 1.94 +1.48 0.36 0.035
TiB42H2 0.38 0.22 2.02 2.02 +1.44 0.34 0.030
TiB41CH4 0.41 2.71 +1.53 0.38 0.029
TiB42CH4 0.32 0.23 2.74 3.61 +1.53 0.37 0.022
Published on 25 May 2019. Downloaded by University of Frankfurt on 7/26/2019 4:59:43 AM.

TiB41H21CH4 0.38 0.22 1.94 3.48 +1.48 0.36 0.030 0.019

between C and Ti was 2.71 Å. When the second CH4 molecule


was added, it was a little far away from the Ti site (dTi–g =
CH4
3.61 Å), whereas the Econ was 0.23 eV per CH4, still showing
the possibility of adsorption. For the mixture adsorption, the
CH4
Econ value of 0.22 eV per CH4 was close to that of pure CH4.
However, considering the energy storage capacity, TiB4–2CH4
was the better choice.
To obtain further information on the adsorption mecha-
nism for the gas molecules on the surface of the TiB4 mono-
layer, we performed a Bader charge analysis.49 It could be seen
from Table 1 that, in the pure TiB4 sheet, each Ti transferred
B1.51 e to the B network; this was consistent with the former
ELF result and confirmed the ionic interactions between Ti and
B. In the gas-adsorbed TiB4 systems, each gas molecule had a
small and negative charge, indicating Coulomb interactions
between the gas molecules and the host materials. When H2
Fig. 2 The side and top views of the optimized TiB4 monolayers with (a) adsorbed onto the TiB4 sheet, it obtained a charge of 0.030 to
zero gas molecule; (b) one H2; (c) two H2; (d) one CH4; (e) two CH4; (f) one
0.035 e and changed the charges of both Ti and B; this
H2 and one CH4. H and C atoms are denoted by white and grey spheres,
respectively.
indicated that the adsorbed H2 interacted with these atoms.
For the CH4 adsorption cases, CH4 had a relatively small charge
( 0.019 to 0.029 e). Moreover, compared to the TiB4 sheet,
H2 were adsorbed, as shown in Fig. 2(c), each H2 molecule had only the charge of Ti changed after the addition of CH4; this
the bond length of 0.78 Å. Fig. 2(f) shows that when two CH4 indicated that the electrostatic interaction in the CH4–TiB4
molecules adsorbed, the first adsorbed CH4 molecule had the system resulted from CH4 and Ti. In addition, considering
same structure as that of the TiB4–CH4 system, whereas the the adsorption energy, we found that it was positively correlated
C–H bond lengths of the second CH4 molecule were consistent with the absolute charge value of the corresponding gas
with that of a free CH4 molecule (1.10 Å). Fig. 2(f) shows molecule.
the optimized TiB4–CH4–H2 system, in which the bond length The density of states (DOS) calculation is another way to
of the adsorbed H2 molecule is 0.79 Å. Compared to the investigate the bond interactions of the TiB4 sheet. As shown in
second adsorbed CH4 molecule in the TiB4–2CH4 system, Fig. 3(a), the DOS peaks for the TiB4 sheet occurred at approxi-
the C–H bond lengths in Fig. 2(f) were the same (1.10 Å), but mately 3 eV, which originated from the interactions between
the H atoms of the CH4 molecule were located in different the B p electrons and the Ti d electrons and were in good
areas: three H atoms were near the TiB4 surface, and the agreement with previously reported results.31 After H2 adsorbed
remaining H atom was off the surface. Using eqn (1) and (2), it on the TiB4 sheet (Fig. 3(b)), it was found that the H 1s orbital
was found that all the adsorption binding energies calculated hybridized with the Ti 3d and B 2p orbital at 7 to 8 eV below
herein lied in the previously reported ideal range of 0.2–0.6 eV the Fermi level. Therefore, the orbital interactions also con-
per molecule.47,48 We found that the binding energy of one tributed to the H2 adsorption. As presented in Fig. 3(c), the
gas molecule was 0.54 eV per H2 (0.41 eV per CH4), and it adsorption of the CH4 molecules onto the TiB4 sheet did not
decreased to 0.38 eV per H2 (0.32 eV per CH4) due to the influence the binding interactions between the Ti and the B
relatively weak interaction between the second gas molecule atoms. Moreover, no orbital interactions existed between the
and the TiB4 sheet, as shown in Table 1. It could be seen CH4 molecules and the host sheet.
from Fig. 2 and Table 1 that the distance between H2 and the To satisfy the DOE target for practical use, it is necessary to
open Ti site was B2.02 Å, much shorter than that for the investigate the adsorption and desorption kinetics of the gas
adsorbed CH4. For one CH4 adsorption case, the distance molecules on the TiB4 monolayer. To illustrate this, we carried

This journal is © the Owner Societies 2019 Phys. Chem. Chem. Phys., 2019, 21, 13151--13156 | 13153
View Article Online

Paper PCCP
Published on 25 May 2019. Downloaded by University of Frankfurt on 7/26/2019 4:59:43 AM.

Fig. 3 Density of states for the compounds (a) TiB4, (b) TiB4–2H2 and (c) TiB4–2CH4. For each compound, total DOS are shown in black at the top of the
figures. The projected DOS on the atoms with s, p and d contributions are shown in red, green and blue, respectively. The origin of energy was set at the
highest occupied state and is indicated with a vertical dashed line.

out ab initio molecular dynamics (AIMD) calculations50 on the of 10.14 wt%, which was comparable to that of previously
optimized 2  2  1 TiB4–2H2 and TiB4–2CH4 super cells at reported Li-decorated g-CN (10.81 wt%)17 and Ti-decorated
different temperatures (50 K, 100 K, 150 K, 200 K, 300 K, 400 K g-C3N4 (9.70 wt%).14 The abovementioned behaviors were consistent
and 500 K) followed by equilibration for 10 000 steps at the 1 fs with the adsorption energies and showed that as energy storage
time scale. During the MD simulation, the gas molecules that materials, TiB4 had great potential for practical applications.
were relatively far away from the surface were deleted. Fig. 4 With regard to the energy storage properties of the TiB4
shows the structures after 10 ps running. It could be seen monolayer system, the distance between adjacent layers was
that the desorption behaviors were the same for TiB4–2H2 and large enough to avoid an interaction. For practical applications,
TiB4–2CH4 at 50 K, 100 K, and 150 K, with zero, one, and three the adjacent TiB4 monolayers may be close enough to each
corresponding gas molecules evaporated, respectively. Five and other to get stacked. Therefore, the multilayer structure of TiB4
four H2 molecules remained adsorbed at 200 K and 300 K. and its corresponding energy storage properties should also be
When the temperature reached 500 K, three H2 molecules were considered. To illustrate this, we investigated the bilayer TiB4
still stably adsorbed, leading to only a 2.65 wt% of reversible system. Several different TiB4 bilayers were optimized, and the
hydrogen storage capacity in the TiB4–2H2 system. As for most stable structure is shown in Fig. 5(a). It could be seen that
TiB4–2CH4, four and two CH4 molecules adsorbed at 200 K each Ti atom in one layer was above the center of the four
and 300 K, and all were desorbed at temperatures Z400 K. membered B ring in the other layer. Moreover, the B–B layer
Considering that the combustion heat for one CH4 molecule distance and interlayer interaction energy were 2.80 Å and
was 3.1 times that of one H2 molecule, the reversible stored energy 329 meV per atom, similar to the previously reported values
of TiB4–2CH4 was equivalent to the gravimetric hydrogen density of 2.78 Å and 351 meV per atom, respectively.31 Via the Bader

Fig. 4 Snapshots after 10 ps MD simulations of TiB4–2H2 and TiB4–2CH4 systems at different temperatures.

13154 | Phys. Chem. Chem. Phys., 2019, 21, 13151--13156 This journal is © the Owner Societies 2019
View Article Online

PCCP Paper

sheet, the interactions were mainly ionic between Ti and B and


covalent in the B network. The stability of TiB4 was demon-
strated by the phonon spectrum calculation. When the H2 and
CH4 gas molecules were adsorbed, they preferred to locate
above the Ti sties, and each side of Ti could adsorb only
one molecule. The first adsorbed molecule had much higher
binding energy than the second adsorbed molecule; this was
consistent with the AIMD results. Via comparison, it was found
that TiB4–2CH4 had better reversibility than TiB4–2H2 and
obtained the effective energy storage of 10.14 wt% under
Published on 25 May 2019. Downloaded by University of Frankfurt on 7/26/2019 4:59:43 AM.

ambient conditions. Moreover, when the TiB4 monolayers were


stacked, their energy capacity dramatically decreased. Thus, in
practical application, the interlayer spacing should be noted.
We expect that this study will provide new insights into
molecular hydrogen storage and accelerate the experimental
studies on energy economy.

Conflicts of interest
Fig. 5 The side and top views of the optimized TiB4 bilayers with (a) zero There are no conflicts to declare.
gas molecules, (b) two H2 molecules, and (c) two CH4 molecules.

charge analysis, it was found that the charges of the Ti and B atoms Acknowledgements
in each layer of the stacked TiB4 were similar to those in the The authors gratefully acknowledge the assistance of Dr Süleyman
individual TiB4 monolayers; this indicated that the moderate Er, Dutch Institute for Fundamental Energy Research in perform-
interlayer interactions were mainly caused by ionic interactions. ing the simulation.
For the energy storage of the TiB4 bilayer, due to the interlayer
interactions and space cases, there was no inside site for gas
adsorption. After optimization, we found that gas molecules tended Notes and references
to adsorb at the Ti sites and above the bilayer surface. For H2
absorption, as shown in Fig. 5(b), each Ti site could only adsorb 1 R. Coontz and B. Hanson, Science, 2004, 305, 957.
one H2 molecule. The adsorbed H2 molecule had the charge of 2 S. Er, G. A. de Wijs and G. Brocks, J. Phys. Chem. Lett., 2010,
0.0061 e, whose absolute value was much smaller than that in the 1, 1982–1986.
TiB4 monolayer systems; this resulted in the weak adsorption 3 C. Zhou, Z. Z. Fang, C. Ren, J. Li and J. Lu, J. Phys. Chem. C,
energy of 0.08 eV per H2 that was unsuitable for mobile applica- 2013, 117, 12973–12980.
tions. Fig. 5(c) shows the optimized configuration of the CH4- 4 H. Yukawa, M. Takagi, A. Teshima and M. Morinaga,
adsorbed TiB4 bilayer. Every Ti site of the TiB4 bilayer also attracted J. Alloys Compd., 2002, 330, 105–109.
a maximum of one CH4 molecule, and the absolute charge of the 5 M. V. Lototsky, V. A. Yartys and I. Y. Zavaliy, J. Alloys Compd.,
adsorbed CH4 molecule ( 0.012 e) was still smaller than that in the 2005, 404, 421–426.
corresponding monolayer cases, whereas it maintained the mod- 6 Z. Liu, L. Xiong, J. Li, S. Liu and S. Er, Int. J. Hydrogen Energy,
erate binding energy of 0.22 eV per CH4. This binding energy could 2018, 43, 21441–21450.
be caused by the charge of the C atom. Via further charge analysis, 7 DOE Technical Targets for Onboard Hydrogen Storage for
we found that the C atom of CH4 in the TiB4 bilayer systems LightDuty Vehicles, http://www.energy.gov/eere/fuelcells/
obtained a negative charge ( 0.13 e), which was in the range of the doe-technical-targets-onboard-hydrogen-storage-light-duty-
C charges for the TiB4 monolayer systems ( 0.11 to 0.17 e). This vehicles, accessed October, 2018.
negatively charged C would interact with a Ti atom (+1.51 e); this 8 A. Zuttel, P. Wenger, S. Rentsch, P. Sudan, P. Mauron and
would lead to effective CH4 adsorption. As a result, the maximum C. Emmenegger, J. Power Sources, 2003, 118, 1–7.
stored energy in the TiB4 bilayer was equivalent to the gravimetric 9 S. Er, G. A. de Wijs and G. Brocks, J. Phys. Chem. C, 2009,
hydrogen density of 5.07 wt%, only half that of the TiB4 monolayer. 113, 18962–18967.
10 S. S. Mao, S. H. Shen and L. J. Guo, Prog. Nat. Sci., 2012, 22,
522–534.
4. Conclusions 11 S. Banerjee, C. G. S. Pillai and C. Majumder, Appl. Phys. Lett.,
2013, 102, 073901.
In summary, we studied the 2D TiB4 monolayer for high energy 12 J. Wang, H. Y. Zhao and Y. Liu, ChemPhysChem, 2014, 15,
density storage by first principles calculations. For the TiB4 3453–3459.

This journal is © the Owner Societies 2019 Phys. Chem. Chem. Phys., 2019, 21, 13151--13156 | 13155
View Article Online

Paper PCCP

13 S. Er, G. A. de Wijs and G. Brocks, J. Mater. Chem. A, 2015, 3, 32 Y. D. Ma, Y. Dai, W. Wei, L. Yu and B. B. Huang, J. Phys.
7710–7714. Chem. A, 2013, 117, 5171–5177.
14 W. Zhang, Z. Zhang, F. Zhang and W. Yang, Appl. Surf. Sci., 33 Q. Deng, T. Wu, G. Chen, H. A. Hansen and T. Vegge, Phys.
2016, 386, 247–254. Chem. Chem. Phys., 2018, 20, 5173–5179.
15 A. Hashmi, M. U. Farooq, I. Khan, J. Son and J. Hong, 34 Y. F. Zhao, M. T. Lusk, A. C. Dillon, M. J. Heben and
J. Mater. Chem. A, 2017, 5, 2821–2828. S. B. Zhang, Nano Lett., 2008, 8, 157–161.
16 X. Tan, H. A. Tahini and S. C. Smith, Energy Storage 35 M. Eddaoudi, J. Kim, N. Rosi, D. Vodak, J. Wachter,
Materials, 2017, 8, 169–183. M. Keeffe and O. M. Yaghi, Science, 2002, 295, 469.
17 Y.-D. Chen, S. Yu, W.-H. Zhao, S.-F. Li and X.-M. Duan, Phys. 36 Q. Xue, M. Wu, X. C. Zeng and P. Jena, J. Mater. Chem. A,
Chem. Chem. Phys., 2018, 20, 13473–13477. 2018, 6, 8916–8922.
Published on 25 May 2019. Downloaded by University of Frankfurt on 7/26/2019 4:59:43 AM.

18 Z. Sheng, S. Wu, X. Dai, T. Zhao and Y. Hao, Phys. Chem. 37 G. Kresse and J. Hafner, Phys. Rev. B: Condens. Matter Mater.
Chem. Phys., 2018, 20, 13903–13908. Phys., 1993, 48, 13115–13118.
19 L. Wang, X. Chen, H. Du, Y. Yuan, H. Qu and M. Zou, Appl. 38 G. Kresse and J. Furthmuller, Comput. Mater. Sci., 1996, 6,
Surf. Sci., 2018, 427, 1030–1037. 15–50.
20 X. Liang, S.-P. Ng, N. Ding and C.-M. L. Wu, Appl. Surf. Sci., 39 G. Kresse and J. Furthmuller, Phys. Rev. B: Condens. Matter
2019, 473, 174–181. Mater. Phys., 1996, 54, 11169–11186.
21 S. S. Han, W.-Q. Deng and W. A. Goddard III, Angew. Chem., 40 P. E. Blochl, Phys. Rev. B: Condens. Matter Mater. Phys., 1994,
Int. Ed., 2007, 46, 6289–6292. 50, 17953–17979.
22 Y. H. Hu and L. Zhang, Adv. Mater., 2010, 22, E117–E130. 41 G. Kresse and D. Joubert, Phys. Rev. B: Condens. Matter
23 E. Tylianakis, E. Klontzas and G. E. Froudakis, Nanoscale, Mater. Phys., 1999, 59, 1758–1775.
2011, 3, 856–869. 42 J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett.,
24 M. Samolia and T. J. D. Kumar, J. Phys. Chem. C, 2014, 118, 1996, 77, 3865–3868.
10859–10866. 43 S. Grimme, J. Antony, S. Ehrlich and H. Krieg, J. Chem. Phys.,
25 Y. Pramudya and J. L. Mendoza-Cortes, J. Am. Chem. Soc., 2010, 132, 154104.
2016, 138, 15204–15213. 44 C. D. Zeinalipour-Yazdi, J. S. J. Hargreaves, S. Laassiri and
26 K. Koizumi, K. Nobusada and M. Boero, Phys. Chem. Chem. C. R. A. Catlow, Phys. Chem. Chem. Phys., 2017, 19,
Phys., 2019, 21, 7756–7764. 11968–11974.
27 Y. W. Li and R. T. Yang, J. Am. Chem. Soc., 2006, 128, 45 F. Bakhshi and N. Farhadian, Int. J. Hydrogen Energy, 2018,
8136–8137. 43, 8355–8364.
28 A. G. Wong-Foy, A. J. Matzger and O. M. Yaghi, J. Am. Chem. 46 A. Togo and I. Tanaka, Scr. Mater., 2015, 108, 1–5.
Soc., 2006, 128, 3494–3495. 47 S. K. Bhatia and A. L. Myers, Langmuir, 2006, 22, 1688–1700.
29 L. J. Murray, M. Dinca and J. R. Long, Chem. Soc. Rev., 2009, 48 Y.-H. Kim, Y. Zhao, A. Williamson, M. J. Heben and
38, 1294–1314. S. B. Zhang, Phys. Rev. Lett., 2006, 96, 016102.
30 S.-Y. Liu, P. Kundu, T.-W. Huang, Y.-J. Chuang, F.-G. Tseng, 49 G. Henkelman, A. Arnaldsson and H. Jonsson, Comput.
Y. Lu, M.-L. Sui and F.-R. Chen, Nano Energy, 2017, 31, Mater. Sci., 2006, 36, 354–360.
218–224. 50 D. Marc and J. Hutter, Ab Initio Molecular Dynamics: Theory
31 X. Qu, J. Yang, Y. Wang, J. Lv, Z. Chen and Y. Ma, Nanoscale, and Implementation, John von Neumann Institute for Com-
2017, 9, 17983–17990. puting, Julich, 2000.

13156 | Phys. Chem. Chem. Phys., 2019, 21, 13151--13156 This journal is © the Owner Societies 2019

You might also like