You are on page 1of 13

Molecular chemisorption on open metal sites in Cu 3 ( benzenetricarboxylate ) 2 : A

spatially periodic density functional theory study


Taku Watanabe and David S. Sholl

Citation: The Journal of Chemical Physics 133, 094509 (2010); doi: 10.1063/1.3479041
View online: http://dx.doi.org/10.1063/1.3479041
View Table of Contents: http://scitation.aip.org/content/aip/journal/jcp/133/9?ver=pdfcov
Published by the AIP Publishing

Articles you may be interested in


Magnetic properties of single 3d transition metals adsorbed on graphene and benzene: A density functional
theory study
J. Appl. Phys. 110, 064303 (2011); 10.1063/1.3636112

Coadsorption of CO and NO on the Cu 2 O ( 111 ) surface: A periodic density functional theory study
J. Chem. Phys. 131, 174503 (2009); 10.1063/1.3251055

Adsorption of transition-metal atoms on boron nitride nanotube: A density-functional study


J. Chem. Phys. 125, 044711 (2006); 10.1063/1.2218841

Metal nanoparticles as models of single crystal surfaces and supported catalysts: Density functional study of size
effects for CO/Pd(111)
J. Chem. Phys. 117, 9887 (2002); 10.1063/1.1516798

Molecular and dissociative chemisorption of NO on palladium and rhodium (100) and (111) surfaces: A density-
functional periodic study
J. Chem. Phys. 108, 6447 (1998); 10.1063/1.476051

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP: 147.143.2.5
On: Mon, 22 Dec 2014 04:35:03
THE JOURNAL OF CHEMICAL PHYSICS 133, 094509 共2010兲

Molecular chemisorption on open metal sites in


Cu3„benzenetricarboxylate…2: A spatially periodic
density functional theory study
Taku Watanabe and David S. Sholla兲
School of Chemical & Biomolecular Engineering, Georgia Institute of Technology,
Atlanta, Georgia 30332-0100, USA
共Received 16 March 2010; accepted 22 July 2010; published online 7 September 2010兲

Chemical interactions of H2O, CO, NO, pyridine, C2H2, H2S, and NH3 with open metal sites in the
metal-organic framework 共MOF兲 Cu3共benzenetricarboxylate兲2 are examined using plane wave
periodic density functional theory 共DFT兲. In the case of single molecule adsorption on a Cu dimer,
NH3 and pyridine have the strongest binding, while NO binds weakly. Binding of pairs of molecules
on a Cu dimer shows significant interaction energies, that is, the binding energy of the pair of
molecules is not a simple summation of the binding energies of each molecule. The effect of
molecular adsorption on the magnetic moments of Cu atoms in the MOF is also examined. Using
the binding energies from DFT calculations, the effects of the pressure and temperature on the
chemisorbed species are investigated. Finally the effect of water adsorption on the elastic behavior
of Cu3共BTC兲2 is described. © 2010 American Institute of Physics. 关doi:10.1063/1.3479041兴

I. INTRODUCTION observed that the as-made sample showed continuous mass


loss of 32% up to 523 K. High temperature x-ray diffraction
Metal-organic frameworks 共MOFs兲 have been exten- 共XRD兲 indicated that the reflection corresponding to the wa-
sively studied for a variety of applications including gas ter molecules bound to Cu sites diminished around 473 K.
separation, storage, and catalysis.1 Some MOFs are known to They also reported that adsorption of gaseous ammonia re-
have metal sites that can be activated by solvent removal and sulted in an irreversible structural change to another crystal,
these sites can play an important role in adsorption but no information was given regarding the structure of the
behavior.2,3 These metal sites create the possibility of allow- final product. Xiao et al.15 performed gravimetric adsorption
ing adsorption of molecules via chemisorption, as distinct and IR spectra measurements of NO and H2 on Cu3共BTC兲2.
from the physisorption that controls that adsorption of mol- The NO adsorption experiments were performed at 196 and
ecules in many MOFs and other nanoporous materials. 296 K, and a significant amount 共⬃2.2 mmol/ g兲 of NO was
Cu3关benzenetricarboxylate共BTC兲兴2, also known as found to remain adsorbed after the reduction of gas pressure.
HKUST-1, is a well-known MOF with open metal sites.4–9 Infrared 共IR兲 spectra at room temperature indicated the for-
Cu3共BTC兲2 has Fm3̄m symmetry and a three dimensional mation of Cu2+ – NO adducts by the growth of a new com-
channel of pores, with adjacent square cavities of approxi- ponent at 1887 cm−1 upon contact with 30 Torr of NO. This
mately 10⫻ 10 Å2 connected by a square window of ⬃7 peak is consistent with the peak observed when NO is ad-
⫻ 7 Å2.4,10,11 In this MOF, Cu dimers are bridged by BTC sorbed in Cu-ZSM-5 although the width is much narrower.16
linkers forming a paddlewheel structure. As synthesized, Prestipino et al.8 described spectroscopic characterization of
these Cu sites are saturated by water molecules that can be Cu3共BTC兲2 upon activation and adsorption of guest mol-
removed by heating the sample at moderate temperature ecules. Interaction of CO and H2 was investigated using
共⬃500 K兲.7,8 After this process, guest molecules can interact x-ray adsorption near edge structure 共XANES兲 and IR spec-
with the open metal sites. Other examples of MOFs that troscopy at 77 K. After the adsorption of CO in dehydrated
exhibit open metal sites include CPO-273,12 and MIL-101.13 Cu3共BTC兲2, a substantial change in the XANES pre-edge
Sorption properties of Cu3共BTC兲2 for a wide variety of feature of Cu2+ species was observed, indicating the modifi-
molecules have been investigated by a number of groups. cation of the first coordination sphere of Cu2+ atoms. The IR
Wang et al.14 reported adsorption of N2, O2, CO, CO2, N2O, spectra were collected by dosing a CO equilibrium pressure
CH4, ethylene, ethane, n-dodecane, and H2O. In their experi- of 1 Torr and by progressively reducing this pressure to
ments, CO uptake was found to be significantly lower 10−3 Torr. This IR study showed a prominent peak at
共⬃0.7 mmol/ g at 1 bar兲 than CO2 共⬃5 mmol/ g兲, C2H4 2178 cm−1 ascribed to Cu2+ – CO adducts that is consistent
共⬃6.2 mmol/ g兲, and N2O 共⬃5.3 mmol/ g兲, and similar to with the same adducts in Cu-ZSM-5.17 As the CO pressure
those of CH4, N2, and O2. Schlichte et al.7 examined the was decreased, the peak shifted downward to 2174 cm−1 and
thermal stability and catalytic properties of Cu3共BTC兲2. They almost completely disappeared at 10−3 Torr. Bordiga et al.18
reported additional IR spectra measurements on Cu3共BTC兲2
a兲
Author to whom correspondence should be addressed. Electronic mail: including CO2 and N2 in addition to CO, NO, and H2.
david.sholl@chbe.gatech.edu. Cu2+ – NO adducts were indicated by the peak at 1887 cm−1

0021-9606/2010/133共9兲/094509/12/$30.00 133, 094509-1 © 2010 American Institute of Physics


This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP: 147.143.2.5
On: Mon, 22 Dec 2014 04:35:03
094509-2 T. Watanabe and D. S. Sholl J. Chem. Phys. 133, 094509 共2010兲

that increases proportionally with the coverage and without materials. Motivated by these observations, we examined
substantial shifts. Clear shifts in the adsorbate vibrational chemisorption of selected molecules on the open Cu site in
spectra were observed for CO 共+36 cm−1兲, CO2 Cu3共BTC兲2 using plane wave DFT.
共−19 cm−1兲, and NO 共+11 cm−1兲 but not for N2 and H2. We We will investigate the chemisorption behavior of H2O,
are not aware of any similar studies that examined gas mix- CO, NO, H2S, NH3, C2H2, and pyridine. H2O is a common
tures. solvent used in the synthesis of many MOFs and separating
Quantum chemistry studies of adsorption of gas mol- H2O from a gas mixture is also useful for producing dry
ecules in MOFs are limited.19 A large fraction of the earlier gases. CO and NH3 are important molecules for industrial
studies focused on H2 adsorption in MOFs because of the applications, but they are also common contaminants for the
strong interest in hydrogen storage.20,21 However, this is one metal catalyst on fuel cell anode materials.28 NO is an im-
of the most challenging systems to study with DFT because portant molecule for biological systems and its delivery is
the interaction is dominated by weak van der Waals forces attractive for antibacterial, antithrombotic, and wound-
where standard DFT is inaccurate. Higher level quantum healing applications.29 H2S is a well-known contaminant in
chemistry approaches such as Møller–Plesset 共MP兲 theory natural gas and corrodes metallic parts used for transporta-
are necessary to describe H2-MOF interactions accurately. tion and processing.30 Pyridine was chosen since it is large
Gao and Zeng22 used DFT and the second-order MP 共MP2兲 enough 共⬃4.9 Å from H to H across the ring兲 to signifi-
methods with correlation consistent triple-zeta basis sets 共cc- cantly modify the channel within Cu3共BTC兲2 but also small
pVTZ and aug-cc-pVTZ兲 to study the adsorption on ben- enough to diffuse through the channels of the framework.
zenelike linker fragments. They showed that the MP2 result Qualitatively, Cu3共BTC兲2 is a Lewis acid,9 so good Lewis
falls between DFT results with the local density approxima- bases such as NH3 and pyridine are expected to bind
tion 共LDA兲 and generalized gradient approximation 共GGA兲, strongly.
and the binding energy can change 5 kJ/mol depending on One possible effect of adsorption of molecules in MOFs
the H2 molecule orientation. Sagara et al.21,23 also performed is the change in their mechanical response. A small number
MP2 calculations with a split-valence quadruple-zeta plus of experiments have already studied the mechanical proper-
polarization 共QZVPP兲 basis set on IRMOF-1 fragments and ties of Cu3共BTC兲2. Allendorf et al.31 used the adsorbate in-
found binding energies of 4–9 kJ/mol on the benzyl linker duced mechanical response of Cu3共BTC兲2 to create a gas
areas and 10 kJ/mol on the ZnO fragment. sensing device. In their work, Cu3共BTC兲2 was deposited on a
Fewer studies have been done using quantum chemistry piezoelectric cantilever and changes in the resistivity were
calculations on adsorption of other gases. Xiang et al.10 per- measured with varying amounts of moisture in the gas and
formed periodic DFT calculations to determine the binding the presence of gas species including methanol, ethanol, and
energies and geometries of adsorption of acetylene in CO2. It was found that the device responded more sensitively
Cu3共BTC兲2. They found that acetylene binds to the frame- to CO2 when Cu3共BTC兲2 is hydrated. Direct measurement of
work along the O–Cu–O bond with binding energies of the elastic property of Cu3共BTC兲2 at high pressure was per-
44.5 kJ/mol with LDA and 15.8 kJ/mol with a GGA func- formed by Chapman et al.32 They measured the relative lat-
tional. Dubbeldam et al.24 investigated the interaction of tice change of Cu3共BTC兲2 as a function of pressure at pres-
small molecules including CO2 and N2 in IRMOF-1 by DFT sures up to 10 GPa using a diamond anvil cell. Fluorinert,
with PBE/ 6-31Gⴱ and classical force fields. They examined isopropanol, and a methanol-ethanol-water 共MEW兲 mixture
the effect of distances and angles on the adsorption energies were used as the fluid media in these experiments. A clear
of the two molecular species and found that the linear mol- change in the pressure dependence of the lattice volume was
ecules preferred to bind along the organic linker. Ramsahye observed when MEW or isopropanol was used, but when no
et al.25 characterized the CO2 adsorption sites in MIL-53 and fluid or fluorinert was used, no such change in the mechani-
MIL-47 using the PW91-GGA functional in periodic simula- cal response was found. These studies involve mainly phys-
tion cells. A number of possible adsorption geometries were isorption of molecules, but little is known about the effect of
identified in the two polymorphs of MIL-53 and adsorption chemisorption on similar properties. Thus, we investigate the
energies were calculated. CPL-1 and CPL-2 have been stud- effect of water adsorption on the elastic property of
ied by Nagaoka et al.26 using periodic LDA-DFT calculation Cu3共BTC兲2 using plane wave DFT.
to characterize the bond length and angles of O2. They re-
ported a binding energy of ⬃40 kJ/ mol. Most of these stud- II. METHODS
ies focus on the adsorption of weakly binding molecules in
the MOFs. We have previously investigated the adsorption As discussed above, we performed calculations that used
and diffusion of CO2 / CH4 in Cu 共hfipbb兲0.5 · 共H2hfipbb兲0.5 by the fully periodic structure of Cu3共BTC兲2. Our computational
combining molecular simulations and DFT.27 This material supercell was one rhombohedral primitive cell of
has the interesting property that CO2 diffuses far faster than Cu3共BTC兲2, which contains 156 framework atoms in its de-
CH4 through its pores. DFT is generally reliable for the for- hydrated form. All calculations with this approach were per-
mation and breaking of chemical bonds. Plane wave DFT is formed with the Vienna ab initio simulation package33 ver-
useful for extended solid materials such as MOFs and this sion 4.6 with the PW91-GGA 共Ref. 34兲 exchange-correlation
approach is far less computationally demanding than higher functional and an energy cutoff of 500 eV. This energy cutoff
level quantum chemistry approaches, so it is capable of mod- was to give well converged total energies for test calculations
eling the large unit cell structures that define many MOF with the crystal structure reported by Chui et al.4 Trial cal-
This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP: 147.143.2.5
On: Mon, 22 Dec 2014 04:35:03
094509-3 Chemisorption on Cu sites in CuBTC J. Chem. Phys. 133, 094509 共2010兲

TABLE I. Crystal structure of dehydrated and hydrated Cu3共BTC兲2 from the


reported experimental values and our DFT optimized structures. All lengths
are in angstrom.

Dehydrated Hydrated

Expt. DFT Expt. DFT

a 26.224a 26.51 26.343b 26.59


Cu–Cu 2.50a 2.48 2.682,b 2.64a 2.60
Cu–O 1.94a 1.97 1.952,b 1.96a 1.99
Cu– OH2 2.165,b 2.19a 2.32
a
Reference 8.
b
Reference 4.

culations were performed sampling k-space at the ⌫-point


and using 2 ⫻ 2 ⫻ 2 k-points. The energy difference between FIG. 1. Molecular structure of a hydrated Cu acetate molecule.
these calculations was less than 3 ⫻ 10−4 eV/ Cu atom. Thus,
all further calculations were performed sampling reciprocal typically placed at the experimentally reported Cu– OH2O dis-
space at the ⌫-point only. All geometry optimizations were
tance 共⬃2.2 Å兲. Geometry optimization was then performed
performed using a conjugate gradient method. The total en-
allowing all atoms to move while keeping the unit cell shape
ergy and ionic force convergence criteria for energy minimi-
and volume fixed. In these calculations, the DFT optimized
zation were 10−4 eV and 3 ⫻ 10−2 eV/ Å, respectively.
unit cell volume for the dehydrated material was used. To
Cu3共BTC兲2 is known to show antiferromagnetic cou-
test if keeping the unit cell fixed was appropriate, a test cal-
pling within each Cu dimer.35 Thus, spin polarization was
culation with no unit cell constraints was performed with a
applied in calculations where the atomic magnetic moments
pair of H2O molecules coordinated to a Cu dimer. The
were averaged by the site projected electron density.36 The
change in the energy and volume of the primitive unit cell
system’s overall magnetic moments was relaxed without
from the DFT calculation that held the cell volume fixed was
constraints on the spin states. Total energy calculations were
only ⬃10−3 eV/ Cu and ⬃4 Å3. Therefore, the rest of the
performed for dehydrated Cu3共BTC兲2 in nonmagnetic, ferro-
adsorption calculations were performed with fixed unit cell
magnetic, and antiferromagnetic states. The lowest energy
at the DFT optimized lattice parameter for computational
state was found to be antiferromagnetic. The total energy of
efficiency.
this spin state was lower than the nonmagnetic state by 0.10
The total binding energy of a single molecule or a pair of
eV/Cu atom.
molecules was defined by
Table I shows a comparison of the fully hydrated and
dehydrated crystal structures from experiments and our DFT Eb = E共CuBTC兲 + ⌺iEi共mol兲 − E共CuBTC + mol兲,
results in which both the unit cell size and shape and the
position of individual atoms were allowed to relax. The ini- where E共CuBTC+ mol兲, E共CuBTC兲, and Ei共mol兲 are the to-
tial geometries of the hydrated and dehydrated structures tal energies of Cu3共BTC兲2 with adsorbate molecules,
were the ones reported by the x-ray powder diffraction study Cu3共BTC兲2 only, and molecule of type i in gas phase, respec-
共XRPD兲 of Chui et al.4 The experimental values in the table tively. i can be either 1 or 2, since our calculations were
include the extended x-ray adsorption fine structure 共EX- performed with either a single molecules or a pair of mol-
AFS兲 study by Prestipino et al.8 in addition to the XRPD ecules.
data of Chui et al. Only the lattice parameter and selected The reliability of plane wave DFT was further examined
bond lengths are shown. The calculated unit cell angles re- by comparing the binding energy of two H2O molecules on a
main at 90.0° after geometry optimization. The unit cell pa- single Cu dimer in Cu3共BTC兲2 and a localized basis set cal-
rameters and bond lengths for the framework atoms are in culation using a copper acetate 共CuAc兲 molecule. CuAc is
good agreement, with deviations of less than 1% between the used as a precursor for the synthesis of Cu3共BTC兲2 and has
DFT and experimental structures. The bond length between the same local coordination for the Cu dimer as in
Cu and H2O is larger by 6%–7% in DFT than in the experi- Cu3共BTC兲2.4,38,39 The structure of a hydrated CuAc molecule
mental data. A careful XRD study of both single crystal and is shown in Fig. 1, while two views of the crystal structure of
powder Cu3共BTC兲2 samples was reported by Wu et al.37 Cu3共BTC兲2 is shown in Fig. 2. We used the PW91 and
Both samples showed negative thermal expansion of less B3LYP functionals with a 6-311+ +Gⴱⴱ basis set in GAUSS-
than 0.3% between 100 and 450 K. Thus we believe the IAN03 to optimize this structure, using broken symmetry to
effect of thermal expansion relative to comparing our DFT account for the antiferromagnetic coupling.40 The bond
data to experimental data is negligible in Cu3共BTC兲2. length and angles of the optimized molecular structure are
In our adsorption studies, either a single molecule or a compared in Table II with observations from neutron
pair of molecules is placed adjacent to a single Cu dimer. diffraction.39 The bond length and angles of the covalent
There are six dimers in a primitive cell, so these calculations bonds in the molecule are in excellent agreement between
correspond to only adsorbing molecules on a small fraction the experiment and both localized basis set calculations. The
of the total metal sites. The atom nearest to the Cu dimer was Cu– OH2 bond, however, is 7% 共8%兲 larger in the B3LYP
This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP: 147.143.2.5
On: Mon, 22 Dec 2014 04:35:03
094509-4 T. Watanabe and D. S. Sholl J. Chem. Phys. 133, 094509 共2010兲

FIG. 2. Crystal structure of dehydrated Cu3共BTC兲2 view along the 关100兴


direction 共left兲. Carbon, hydrogen, copper, and oxygen atoms are repre-
sented by the blue, white, brown, and red spheres. The paddlewheel struc-
ture of a Cu dimer in Cu3共BTC兲2 with water molecules coordinated to the
metal centers is shown on the right.

共PW91兲 calculation, an outcome very similar to our plane


wave DFT calculation for the crystalline MOF. Several of the FIG. 3. Binding energies of adsorbate molecules on a Cu dimer. Bars with
angles in the hydrated molecule also deviate from the neu- black, white, and gray are for the binding modes of single molecules, two
tron diffraction data. This may be associated with the fact identical molecules, and combinations, respectively.
that the experimental data are for a molecular crystal of hy-
drated CuAc rather than an isolated molecule. Finally, the
III. RESULTS
binding energy of the pair of water molecules was calculated
to be 0.84 共0.85 and 0.89兲 eV from plane wave DFT with the A. Binding energies
PW91 functional for the MOF 共localized basis set calcula- To illustrate our approach to examining the binding of
tions with the B3LYP and PW91 functionals for the CuAc H2O, CO, NO, H2S, C2H2, NH3, and pyridine, we first con-
molecule兲, showing good agreement between the calcula-
sider CO inside the pores of Cu3共BTC兲2. We first performed
tions.
calculations in which the center of a CO molecule was
The majority of our calculations examined molecules
placed in the center of the MOF main cavity with a random
binding to a single dimer in a Cu3共BTC兲2 unit cell. To probe
orientation and then the geometry of the entire system was
the importance of interactions among separate dimers on mo-
optimized as described above. This calculation gave a bind-
lecular adsorption energies, we calculated the adsorption en-
ing energy of only 0.003 eV and no obvious change in the
ergy for each pair of H2O molecules in fully hydrated
structure of Cu3共BTC兲2 or the bond length of CO was de-
Cu3共BTC兲2. This energy was 0.82 eV, which differs only
tected. This indicates, unsurprisingly, that configurations ex-
slightly from the 0.84 eV binding energy calculated when
ist for this small molecule inside the MOF where chemical
H2O bound only to a single dimer. This suggested that the
interactions with the MOF framework are negligible. We
latter calculation is sufficient to describe binding of small
molecules to the Cu sites in Cu3共BTC兲2. note the well-known fact that DFT calculations do not cap-
ture dispersion interactions accurately,41 so the contributions
TABLE II. Bond length and angles of a hydrated CuAc molecule from to the molecular adsorption energy in this and all of the
neutron diffraction 共Ref. 39兲 and localized basis set calculations. The labels configurations we consider do not include these interactions.
of atoms for the bond angles are shown in Fig. 1. X – Oi is the average of the
Figure 3 shows the binding energy of all the adsorbate
distances between atom X and O1, O2, O3, and O4.
molecules and their combinations we examined. When inter-
Expt. B3LYP PW91 action of a single molecule with a Cu dimer is considered,
NH3 and pyridine show the highest values of Eb
Bond length 共Å兲
共⬃0.8 eV兲. This is not surprising, since NH3 and pyridine
Cu–Cu 2.614 2.61 2.53
are strong Lewis bases and Cu in Cu3共BTC兲2 is known to be
Cu– OH2 2.151 2.30 2.32
Cu– Oi 1.992 1.97 1.99 a strong Lewis acid.9 Pyridine and bipyridine are important
C – Oi 1.263 1.27 1.27 linker molecules in MOFs and molecular crystals, so it is
C–C 1.500 1.51 1.52 natural that they show strong binding with the metal centers.
Bond angles 共deg兲 CO, H2O, and H2S have similar binding energies of 0.48,
O1-Cu-O2 89.21 89.47 89.69 0.49, and 0.53 eV, respectively. The weakest binding species
O2-Cu-O3 91.05 90.55 90.52 was NO with 0.30 eV. The binding energy of C2H2 was
Cu-O1-C 121.42 123.2 120.33 found to be 0.38 eV. An earlier theoretical study of acetylene
O1-Cu-Ow 93.14 84.21 81.68 binding on Cu sites by Xiang et al. gave binding energies of
O2-Cu-Ow 91.04 82.97 81.76
0.17 共0.46兲 eV with the PBE-GGA 共LDA兲 functional.10 Our
O3-Cu-Ow 98.12 108.45 105.53
O4-Cu-Ow 93.5 104.33 105.78
value lies in between their reported values, but it is not clear
Cu-Ow-Hw 115.01 97.17 94.23 whether these earlier calculations considered the antiferro-
magnetic nature of the Cu dimers.
This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP: 147.143.2.5
On: Mon, 22 Dec 2014 04:35:03
094509-5 Chemisorption on Cu sites in CuBTC J. Chem. Phys. 133, 094509 共2010兲

TABLE III. Pair interaction energies from the DFT calculations. showed that only a single NO was bound to the Cu dimer;
the other NO molecule was displaced to the pore of the cage.
⌬Eb,AA 共eV兲
This result agrees with the adsorption experiment by Xiao et
2 C 2H 2 0.28 al.15 who observed that approximately 2.2 mmol/g of NO
2 NO 0.31 remained strongly bound and was not released from the
2 CO 0.40 sample after heating to 196 or 296 K. This amount is very
2 H 2S 0.32 close to one NO molecule per Cu dimer.
2 H 2O 0.20
2 NH3 0.26
2 pyridine 0.31 B. Structure analysis

⌬Eb,AB 共eV兲 We now turn to the geometry of adsorbates on the metal


CO+ NO 0.23 atoms. The bond lengths and angles between adsorbed mol-
C 2H 2 + H 2O 0.19 ecules and Cu atoms and the separation of Cu atoms within
NO+ H2O 0.07 the Cu dimer are summarized in Table IV. The Cu to mol-
CO+ H2O 0.22 ecule bond length is defined relative to the atom in the mol-
H 2S + H 2O 0.19 ecule that is closest to Cu, which is O for H2O, C for CO and
NH3 + H2O 0.23
C2H2, N for NO, pyridine, and NH3, and S for H2S. For
pyridine+ H2O 0.20
C2H2, the C–Cu distance is averaged between the two C–Cu
bonds. To characterize the adsorbate angle, we define a mo-
lecular axis for each molecule. For CO, NO, and C2H2, this
The binding energies of a single molecule are of limited axis is simply the C–O, N–O, or C–C bond. For H2O, H2S,
use for comparing with available experimental data since the and NH3, the vector is defined by the sum of each O–H,
Cu dimers in the MOF are expected to be saturated in most S–H, or N–H bonds. For pyridine, the axis is from N to the C
situations. Thus, we considered the cases where two mol- on the opposite side of the aromatic ring. The adsorbate
ecules are bound to a Cu dimer. As in the case of single angle is then defined by the angle between the Cu-adsorbate
molecule binding, calculations were initialized by placing bond and the adsorbate’s molecular axis. The Cu atoms in
two molecules ⬃2.2 Å away from the Cu atoms in a dimer. the dimer are distinguished by labeling them as Cu1 and
A principal result of these calculations is that the binding Cu2.
energy of a pair of molecules on a Cu dimer is not simply It is clear from Table IV that the Cu-adsorbate bond
twice the binding energy of the individual molecule. This lengths simply reflect the size of the atoms in the adsorbate
effect can be quantified for molecules A and B binding to a molecule except for C2H2, where two atoms bind to the
Cu dimer by defining the pair interaction energy metal center. H2S and H2O give the largest and second larg-
⌬Eb,AB = Eb,A + Eb,B + Eb,A+B . est values, respectively, as expected from the size of S and O
atoms. NH3 and pyridine are the same distance from the
Here, the terms on the right hand side correspond to the metal center and are aligned almost perfectly along the axis
binding energies of a single molecule of A, a single molecule of the Cu dimer. The rest of the molecules bind to Cu atoms
of B, and the two molecules together. With this definition, a at oblique angles. These angles decrease in the order of CO
positive value of ⌬Eb,AB indicates having both molecules 共162°–165°兲, H2O 共123°–131°兲, NO 共123°–128°兲, and H2S
present on the same Cu dimer gives a smaller net binding 共96°–108°兲. H2S lies almost parallel to the plane of Cu–O
energy than the two molecules would give individually. The bonds of the framework. When H2O 共H2S兲 is present, one of
value of this pair interaction energy for each combination of the two O–H 共S–H兲 bonds point toward the nearest O atom
molecules we considered is summarized in Table III. All the bonded to the Cu atom as a part of the framework. This
interaction energies lie between 0.19 and 0.40 eV, except for orientation indicates formation of a very weak hydrogen
the case of NO+ H2O, where the value is only 0.07 eV. Thus, bond, although the distance between the framework O atom
binding energies of a second molecule on a Cu dimer are and H in the adsorbate is large 共⬃2.8 Å兲 compared to typi-
always lower than the first molecule. cal hydrogen bonds 共⬍2.3 Å兲. In the case of C2H2, the mol-
Comparing the magnitudes of the binding energies, we ecule lies parallel to one of the O–Cu–O bonds along the
can conclude that if two CO, two H2S, two NO, or two C2H2 diagonal direction in the CuO4 plane. Thus unlike other mol-
are adsorbed on a Cu dimer, they can be replaced by two ecules we considered, all the atoms in C2H2 contribute to the
H2O molecules if a sufficient amount of water vapor is avail- binding to the framework. This geometry is also observed by
able. In order to make more careful comparisons with experi- Xiang et al.10 although that work did not report bond lengths
mental conditions, we consider the effect of temperature and or angles that would allow a quantitative comparison be-
pressure in Sec. III D. tween our calculations and their results.
It is interesting to note that the binding energy of two Finally, the Cu1-Cu2 distance is larger than the dehy-
NO molecules is almost identical to that of a single NO drated MOF in all cases. This distance varies from 2.51 Å for
molecule. With our definition of the interaction energy, this one NH3 to 2.67 Å for two CO. When CO or pyridine is
means that the interaction energy of two NO molecules is bound to the dimer, the Cu1-Cu2 distance is consistently
equal to the binding energy of a single NO molecule. Exam- larger than the other cases. This result does not appear to be
ining the optimized geometry of two adsorbed NO molecules correlated in a simple way to the molecular binding energies.
This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP: 147.143.2.5
On: Mon, 22 Dec 2014 04:35:03
094509-6 T. Watanabe and D. S. Sholl J. Chem. Phys. 133, 094509 共2010兲

TABLE IV. Bond lengths and angles between Cu atoms and adsorbate molecules, and Cu atom separation
within the dimer. For the mixed adsorbate, A + B, A and B correspond to mol1 and mol2, respectively. Cu1-Cu2
distance in the dehydrated state is 2.48 Å as shown in Table I.

Cu1-mol1 Cu2-mol2

Length Angle Length Angle Cu1-Cu2


Adsorbate 共Å兲 共deg兲 共Å兲 共deg兲 共Å兲

C 2H 2 2.475 90.0 ¯ ¯ 2.517


NO 1.925 127.2 ¯ ¯ 2.548
CO 2.061 165.2 ¯ ¯ 2.590
H 2O 2.304 131.4 ¯ ¯ 2.522
H 2S 2.636 107.7 ¯ ¯ 2.526
NH3 2.171 178.0 ¯ ¯ 2.505
pyridine 2.174 178.6 ¯ ¯ 2.547
2 C 2H 2 2.561 90.0 2.557 89.9 2.565
2 NO 1.972 129.9 ¯ ¯ 2.551
2 CO 2.189 161.9 2.195 165.1 2.666
2 H 2O 2.336 128.8 2.333 124.5 2.547
2 H 2S 2.685 106.3 2.650 95.6 2.584
2 NH3 2.207 177.6 2.209 174.1 2.547
2 pyridine 2.198 178.1 2.173 177.1 2.620
CO+ NO 2.108 165.3 2.079 123.2 2.602
C 2H 2 + H 2O 2.603 90.4 2.395 128.9 2.562
NO+ H2O 1.998 124.8 2.361 128.4 2.553
CO+ H2O 2.013 124.8 2.379 128.4 2.573
H 2S + H 2O 2.658 97.5 2.306 125.7 2.564
NH3 + H2O 2.219 177.4 2.302 122.6 2.547
Py+ H2O 2.163 178.8 2.320 123.1 2.578

C. Magnetization of Cu atoms The magnetization of the Cu atoms in a dimer and the


As mentioned in Sec. II, each Cu dimer in Cu3共BTC兲2 simulation cell when two identical molecules are coordinated
has antiferromagnetic coupling of the magnetic moments and with the dimer is shown in Table VI. Essentially the same
long range ferromagnetic order between the dimers.35 The trend is observed as for NO, C2H2, H2O, and H2S adsorp-
adsorption of CO or NO is known to quench the magnetic tion. NO reduces the magnetization by 55% for Cu1 and 11%
moment of surrounding metal atoms on a metal surfaces such for Cu2. C2H2, H2O, and H2S are again found to have no
as Ni共110兲.42 Our DFT calculations allow us to examine effect on the magnetization. NH3 and pyridine increased the
whether similar effects occur in Cu3共BTC兲2. magnetization by 6 共6兲% for Cu1 and 6 共5兲% for Cu2. The
The magnetization of Cu atoms and the simulation cell only notable difference is that the pair of CO molecules com-
when a single adsorbate molecule is coordinated with a Cu pletely quenched the magnetization of the dimer. The total
dimer are shown in Table V. The magnetization with no ad- magnetization was mostly negligible although a small
sorbate molecules present is also included. All molecules in change of 0.071 Bohr was observed when a pair of NO was
these calculations are bound to Cu1. A range of outcomes present.
can be seen. When a single NO 共CO兲 is present, the magne- We also examined the magnetization of Cu dimers and
tization of Cu1 and Cu2 are reduced by 54 共13兲% and 15 the simulation cell when pairs of unlike molecules were co-
共12兲%, respectively. Adsorption of C2H2, H2O, or H2S ordinated to the dimer. Results from these calculations are
hardly alters the magnetization, but NH3 or pyridine in- summarized in Table VII. In this table, the combination A
creases it by 6%–7%. There is no obvious correlation be- + B refers to a dimer where molecule A 共B兲 is bound to Cu1
tween the binding energies of molecules on the Cu sites and 共Cu2兲. The trend is, in general, consistent with the previous
the impact of the adsorbing molecules on the magnetization results for the adsorption of single and identical pairs of mol-
of the Cu atoms. The total magnetization of the entire simu- ecules. One difference is that when NO 共CO兲 is paired with
lation cell is nearly zero as expected from the collinear anti- H2O, the magnetization of the Cu atom bound to the H2O
ferromagnetic order. also decreased by 7 共5兲%. Similarly to the case of the pair of
TABLE V. Magnetization of Cu atoms of the dimer and the simulation cell with no adsorbate and single
adsorbate molecules. All values are in Bohr.

None C 2H 2 NO CO H 2S H 2O NH3 Pyridine

Cu1 0.488 0.497 0.224 0.426 0.497 0.509 0.523 0.523


Cu2 ⫺0.488 ⫺0.481 ⫺0.415 ⫺0.430 ⫺0.476 ⫺0.483 ⫺0.481 ⫺0.467
Total ⫺0.007 ⫺0.006 0.041 ⫺0.012 ⫺0.015 ⫺0.017 0.032 ⫺0.046

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP: 147.143.2.5
On: Mon, 22 Dec 2014 04:35:03
094509-7 Chemisorption on Cu sites in CuBTC J. Chem. Phys. 133, 094509 共2010兲

TABLE VI. Magnetization of Cu atoms of the dimer and the simulation cell with two identical adsorbate
molecules. All values are in the Bohr.

2 C 2H 2 2 NO 2 CO 2 H 2S 2 H 2O 2 NH3 2 pyridine

Cu1 0.501 0.221 ⫺0.003 0.494 0.500 0.518 0.518


Cu2 ⫺0.498 ⫺0.420 0.005 ⫺0.496 ⫺0.500 ⫺0.517 ⫺0.509
Total 0.001 ⫺0.079 0.000 0.001 0.001 0.000 ⫺0.014

identical molecules, the total magnetization was generally fraction 共0.5–1.0兲 of the adsorption sites, although H2O satu-
small and the highest value was observed when NO and H2O rates the sites at much lower pressure of ⬃5 ⫻ 10−3 bar. As
molecules were bound to a Cu dimer. mentioned in the Introduction, it is known experimentally
that at room temperature Cu3共BTC兲2 is normally saturated
D. Adsorption behavior of Cu3„BTC…2 with H2O in air.4,7,8 Our calculations predict that at least
down to a water partial pressure of 10−3 bar, 90% of the
Knowledge of binding energies of molecules on Cu sites sites are occupied by water molecules. This means that when
is a useful way to qualitatively order the binding strengths of the relative humidity of room temperature air is only ⬃2%,
different molecules, but to make direct predictions about ex- most of the Cu sites are already saturated.
periments, the amount of adsorption for each species under The coverage of CO remains constant at 0.5 for almost
specific conditions must be determined. To this end, we per- the entire range of the pressure, but NO shows a significant
formed grand canonical Monte Carlo 共GCMC兲 simulations drop below 10−3 bar. If a sufficiently wide pressure range is
of adsorption on the Cu sites in Cu3共BTC兲2. GCMC simula- considered, all of the molecules except NO show a two step
tion has been widely used to predict adsorption isotherms in adsorption profile during single component adsorption. For
MOFs in situations where the process occurs via physisorp-
NO, the maximum adsorbed amount within this description
tion and can be modeled using a classical forcefield.6,19,43 In
is one molecule per dimer, as discussed above. When the
our calculations, we take a somewhat different approach by
temperature is raised to 500 K, the isotherm for each species
restricting our attention to molecules that bind to Cu sites in
is shifted towards higher pressures. No molecule can saturate
the geometries characterized by the DFT calculations above.
the adsorption sites below 5 ⫻ 10−3 bar. Our calculation
In this case, Cu3共BTC兲2 defines a spatially regular lattice of
shows that even at this high temperature there is still a water
adsorption sites, and our DFT calculations provide well-
coverage of 0.2–0.5 for H2O partial pressures between 10−3
defined binding energies. We performed calculations using
4 ⫻ 4 ⫻ 4 unit cells, which contained 3072 Cu sites. When and 10−2 bar, and the value falls to nearly zero by 10−4 bar.
two molecules are bound to a single Cu dimer, the pair in- Dehydration of Cu3共BTC兲2 has been investigated by several
teraction energies from the DFT calculations are used to cor- experimental groups and it is known to happen between 450
rect the binding energy of the second molecule. GCMC cal- and 550 K.7,8 The EXAFS study by Prestipino et al.8 sug-
culations were performed with standard methods, treating gests that a small fraction of Cu sites may still be coordi-
each gas phase species as an ideal gas.44 It would be possible nated with H2O even after the dehydration at 453 K. Our
to extend these calculations to include bound and phys- calculations are qualitatively consistent with these observa-
isorbed molecules simultaneously if a forcefield was used to tions.
define the energies of physisorbed molecules, but we have Our GCMC simulations make it straightforward to pre-
not pursued this direction in the current work. dict the composition of molecules bound to Cu sites in equi-
Single component isotherms of NO, CO, H2S, H2O, librium with multicomponent gas phase mixtures. We calcu-
NH3, and pyridine at 300 and 500 K are shown in Fig. 4. The lated the coverage of NO– H2O, CO– H2O, C2H2 – H2O,
partial pressures are set between 10−5 and 1 bar for NO, CO, H2S – H2O, NH3 – H2O, pyridine-H2O, and CO–NO mixtures
H2S, and NH3. The maximum pressure for H2O and pyridine at 300 and 500 K using mixture GCMC simulations. The
is limited to 10−2 bar since their saturation pressures at 300 coverage of NO in NO– H2O and NH3 in NH3 – H2O mix-
K are roughly 0.05 bar for H2O and 0.03 bar for pyridine.45 tures at 300 K are shown in Fig. 5; these two examples
Each isotherm is shown in terms of the fractional coverage of illustrate the behavior that occurs for most of the mixtures
each Cu site. At 300 K, NH3 and pyridine completely satu- we considered. In the NO– H2O mixture, not all of the ad-
rate the metal sites over the entire pressure range because of sorption sites are occupied by adsorbate molecules in the low
their large binding energies. H2O and H2S also cover a large pressure region. When the partial pressures of both NO and

TABLE VII. Magnetization of Cu atoms of the dimer and the simulation cell with mixed adsorbate molecules.
All values are in Bohr.

CO+ NO C 2H 2 + H 2O NO+ H2O CO+ H2O H 2S + H 2O NH3 + H2O Py+ H2O

Cu1 0.324 0.504 0.261 0.431 0.501 0.525 0.541


Cu2 0.389 ⫺0.510 ⫺0.454 ⫺0.463 ⫺0.495 ⫺0.493 ⫺0.483
Total 0.020 0.002 ⫺0.071 0.012 0.018 ⫺0.037 ⫺0.029

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP: 147.143.2.5
On: Mon, 22 Dec 2014 04:35:03
094509-8 T. Watanabe and D. S. Sholl J. Chem. Phys. 133, 094509 共2010兲

FIG. 4. Calculated single component


adsorption isotherms of NO, CO,
C2H2, H2S, H2O, NH3, and pyridine at
300 K 共left兲 and 550 K 共right兲.

H2O are low, there are sufficient open metal sites to which CO–NO and CO–CO pairs, resulting in a change in the trend
each adsorbate can bind so that the coverage of the NO is of the coverage of the metal sites. Complete coverage of the
nearly independent of the partial pressure of H2O. As the metal sites is reached only at the highest pressures we con-
partial pressure of each species increases, all the sites are sidered.
filled and there is competition between NO and H2O. How- At 500 K, a large fraction of the sites is not filled at low
ever, because of the higher binding energy of H2O, H2O pressures in the mixture isotherms we examined. Figure 7
dominates the coverage over NO molecules at the same par- shows the coverages of H2S, H2O, and NH3 in the H2S – H2O
tial pressure. The same trend is observed in C2H2 – H2O, and NH3 – H2O mixtures. The coverages of NO– H2O,
CO– H2O, and H2S – H2O mixtures although the entire iso- CO– H2O, and C2H2 – H2O are qualitatively the same as that
therm is shifted because of the difference in the binding en- of H2S – H2O. The maximum coverage of the adsorbate mol-
ergies. In the NH3 – H2O gas mixture, the coverage of NH3 is ecules in the H2S – H2O mixture is about 0.65, even at the
independent of the partial pressure of NH3 in this pressure highest pressure we examined. The coverage in the
range because of the large difference in the binding energies NH3 – H2O mixture is essentially the coverage of the single
of NH3 and H2O. Even though the binding energy of a pair component NH3, and H2O can occupy only a small fraction
of NH3 is higher than that of either a pair of H2O or of the coverage even at highest H2O partial pressure exam-
NH3 – H2O pair, a significant amount of H2O molecules bind ined. The same qualitative trend as H2S – H2O was also ob-
at high water partial pressures. This indicates that the chemi- served in the CO–NO gas mixture as shown in Fig. 8.
cal potential of H2O is large enough in the range from 10−3
to 10−2 bar to partially displace the strongly bound NH3
molecules. The coverage of pyridine in the pyridine-H2O
mixture is almost indistinguishable from that of NH3 in the
E. Effect of H2O adsorption on the elastic behavior
NH3 – H2O mixture.
of Cu3„BTC…2
A qualitatively different type of behavior occurs for the
CO–NO mixture. The coverages of CO and NO in this mix- In the previous sections, we have focused on the inter-
ture at 300 K are shown in Fig. 6. In this case, not all of the action of small molecules with the open metal sites in
adsorption sites are filled by the adsorbate molecules. In the Cu3共BTC兲2 under equilibrium conditions. As introduced in
low pressure region, there is competition between CO and Sec. I, the mechanical behavior of Cu3共BTC兲2 is known to be
NO and only one molecule is bound to each dimer. As the affected by the adsorption of molecules. Because the poten-
partial pressures increase, the competition becomes between tial effect of chemically bound species on this behavior is not

FIG. 5. Calculated coverages of NO in


NO– H2O binary mixtures 共left兲 and
NH3 in NH3 – H2O 共right兲 binary mix-
tures at 300 K.

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP: 147.143.2.5
On: Mon, 22 Dec 2014 04:35:03
094509-9 Chemisorption on Cu sites in CuBTC J. Chem. Phys. 133, 094509 共2010兲

FIG. 6. Calculated coverages of CO


共left兲 and NO 共right兲 in the CO–NO
binary mixtures at 300 K.

well understood, we briefly consider how the presence of We examined the elastic constants of Cu3共BTC兲2 in both
chemisorbed H2O ligands affects the elastic properties of this its dehydrated and completely hydrated state using plane
MOF. wave DFT. For the hydrated sample, every Cu atom is coor-
Elements of an elastic constant tensor of a crystalline dinated with a H2O molecule. Our rhombohedral primitive
solid are related to the element i of the stress, ␴i, and strain, cell is aligned so that the corresponding cubic unit cell edges
␧i, tensors by Hooke’s law, are along the x, y, and z axes to simplify the calculation. As
␴i = Cij␧j , a result, C11 and C12 are determined by applying uniaxial
strain in the range of ⫺0.015 and 0.015 and calculating the
where i and j run from 1 to 6. With this notation, the elastic
stress ␴1 is the uniaxial stress in the x direction on the plane uniaxial stress in the x and y directions. C44 is obtained by
perpendicular to the x axis, ␴xx. Similarly, index values of applying the shear strain in the x-y plane along the x direc-
共1, 2, 3, 4, 5, and 6兲 correspond to 共xx, yy, zz, xy, yz, and zx兲, tion and calculating the corresponding shear stress. Although
respectively. For the materials with cubic symmetry such as the simulation cell is fixed at each strain, all the atom posi-
Cu3共BTC兲2, there are only three independent nonzero elastic tions are allowed to optimize to find the minimum energy
constants, C11, C12, and C44. configuration.

FIG. 7. Calculated coverages of H2S


共top left兲 and H2O 共top right兲 in
H2S – H2O binary mixtures, and NH3
共bottom left兲 and H2O 共bottom right兲
in NH3 – H2O binary mixtures at
500 K.

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP: 147.143.2.5
On: Mon, 22 Dec 2014 04:35:03
094509-10 T. Watanabe and D. S. Sholl J. Chem. Phys. 133, 094509 共2010兲

FIG. 8. Calculated coverages of CO


共left兲 and NO 共right兲 in CO–NO binary
mixtures at 500 K.

Figure 9 shows the stress tensors as functions of strains molecules on the Cu sites on Cu3共BTC兲2. Based on the DFT
for both dehydrated and hydrated Cu3共BTC兲2. The elastic results, we further investigated the effects of temperature,
constants are obtained from a linear fit of the data to Hooke’s pressure, and presence of other molecules using GCMC
law and the values are shown in Table VIII. C11, C12, and C44 simulations.
for dehydrated 共hydrated兲 Cu3共BTC兲2 are 31.8 共29.2兲, 21.2
共21.5兲, and 12.7 共14.7兲, respectively, with all values in GPa.
Thus, when the MOF is completely hydrated, C11 decreases
by 8.2%, but C12 and C44 increase by 1.4% and 15.7%. These
values of the elastic constants are similar to calculated values
for MOF-5 共C11 = 29.2 GPa, C12 = 13.1 GPa, and C44
= 1.4 GPa by Samanta et al.46 using LDA-DFT, C11
= 27.8 GPa, C12 = 10.6 GPa, and C44 = 3.6 GPa by Bahr
et al. using GGA-DFT, and E = 14.9 GPa by Greathouse and
Allendorf using a classical model兲.51 The reported shear
modulus for MOF-5 is unusually small, but it is not seen in
our calculations for Cu3共BTC兲2.
The bulk modulus of a cubic crystal is 共C11 + 2C12兲 / 3.
Our calculations give a bulk modulus for dehydrated 共hy-
drated兲 Cu3共BTC兲2 of 24.7 GPa 共24.0 GPa兲. The small de-
crease 共2.8%兲 in the bulk modulus upon hydrating the
sample indicates that the MOF becomes slightly softer under
isostatic strain. At the same time, it becomes harder to shear,
as indicated by the increase in C44. We separately calculated
the bulk modulus of dehydrated Cu3共BTC兲2 by applying iso-
static strain. The simulation cell was strained from +1 to
⫺1% in each direction along the unit cell parameters and the
total energies were calculated by DFT. The total energies
were fitted to the Birch–Murnaghan equation of state.47 This
approach gives B = 24.4 GPa, which is in good agreement
with the value obtained from the uniaxial strain. Our values
of the bulk modulus are lower than the experimental value
共⬃30.7 GPa兲 reported by Chapman et al.,32 but the bulk
modulus reported in these experiments showed quite large
variations as the contacting fluid was varied. For example, a
sample with isopropanol shows 25.9 GPa, while MEW can
give as high as 41.9 GPa. It is of course possible that these
diamond anvil cell experiments also measure effects due to
grain interactions or other defects that cannot be probed by
our calculations with a defect-free crystal.

IV. DISCUSSION

We performed DFT calculations to determine the trend FIG. 9. Calculated uniaxial and shear stress-strain from dehydrated 共filled
in the strength of bonds and geometry of the selected small circles兲 and hydrated 共open circles兲 Cu3共BTC兲2.
This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP: 147.143.2.5
On: Mon, 22 Dec 2014 04:35:03
094509-11 Chemisorption on Cu sites in CuBTC J. Chem. Phys. 133, 094509 共2010兲

TABLE VIII. Elements of elastic constants, bulk modulus, and Young’s modulus of Cu3共BTC兲2 determined by
uniaxial and shear strain using plane wave DFT.

C11 C12 C44 B E


共GPa兲 共GPa兲 共GPa兲 共GPa兲 共GPa兲

Dehydrated 31.8 21.2 12.7 24.7 14.9


Hydrated 29.2 21.5 14.7 24.0 11.0

One of our conclusions is that CO bonds with the Cu2+ Cu3共BTC兲2 may be useful for functionalization of
metal center and it can occupy some fraction of Cu sites even Cu3共BTC兲2 since pyridine is relatively large compared to the
in the presence of some water. The IR study by Bordiga et al. size of the cage of the MOF. By controlling the available
clearly shows binding of CO on unsaturated Cu sites.18 How- pore volume, it may be possible to control the adsorption or
ever, there is no other experimental confirmation of strong separation behavior of specific gas mixtures.
binding of CO in Cu3共BTC兲2. Adsorption experiments by The elastic constants of Cu3共BTC兲2 were found to be
Wang et al.14 with CO show a typical type I isotherm with similar to those of IRMOF-1 previously reported by Samanta
weak physisorption. Our results are inconsistent with this et al.46 Although there is no direct measurement of the elastic
experimental observation. One possible explanation for this constants for single crystal Cu3共BTC兲2 available, the small
is that, in their adsorption experiment, H2O molecules were change in the bulk modulus is in line with the experimentally
not completely removed and kinetic limitations keep the CO observed change in the unit cell volume.
molecules from binding to the metal sites. Liu et al.48 Our approach of using DFT and GCMC is not specific to
pointed out that the pore volume from the sample of Wang Cu3共BTC兲2 and should be useful for describing chemisorp-
et al. is much smaller than theoretical values for an ideal tion of molecules in MOFs in general. This type of DFT
crystal and experimental values from carefully prepared calculation should be applied with the caveat that it is not
samples. capable of accurately modeling the interaction of molecules
To the best of our knowledge, no one has explored the with MOFs through van der Waals forces. We have shown
possibility of mixed binding modes. In the case of NO ad- how binding energies obtained from DFT can be used within
sorption, we predict that a small fraction of NO was able to GCMC simulations to give information about the coverages
adsorb on the metal sites even in the presence of water. Even of chemically bound species as functions of temperature and
without the presence of the water, NO was able to cover only partial pressure.
half of all the adsorption sites. This result is consistent with
the experimental adsorption study by Xiao et al.15 ACKNOWLEDGMENTS
Among all the molecules investigated, NH3 and pyridine
showed the highest binding energies. Unfortunately, there is We thank Dr. Thomas Manz for his assistance with the
no experimental report on adsorption of NH3 in Cu3共BTC兲2 B3LYP calculations and Professor Krista Walton for helpful
to which we can compare. Schlichte et al.7 briefly mentioned discussion.
in their report that they observed an irreversible transforma- 1
tion when NH3 is introduced in Cu3共BTC兲2, however, there J. R. Li, R. J. Kuppler, and H. C. Zhou, Chem. Soc. Rev. 38, 1477
共2009兲.
is no further information on the final structure. DFT can be 2
C. Janiak, Dalton Trans. 14, 2781 共2003兲; S. Kitagawa, R. Kitaura, and
used to find a ground state geometry near the starting con- S. Noro, Angew. Chem., Int. Ed. 43, 2334 共2004兲.
3
figuration, but it is not capable of predicting complete struc- J. Lee, O. K. Farha, J. Roberts, K. A. Scheidt, S. T. Nguyen, and J. T.
Hupp, Chem. Soc. Rev. 38, 1450 共2009兲.
tural transformations. We performed a preliminary calcula- 4
S. S.-Y. Chui, S. M.-F. Lo, J. P. H. Charmant, A. G. Orpen, and I. D.
tion by covering all binding sites with NH3 molecules. All Williams, Inorg. Chem. Commun. 283, 1148 共1999兲.
5
the simulation cell parameters and atom positions are al- J. L. C. Rowsell and O. M. Yaghi, J. Am. Chem. Soc. 128, 1304 共2006兲.
6
lowed to relax using the same convergence criteria as the J. R. Karra and K. S. Walton, Langmuir 24, 8620 共2008兲.
7
K. Schlichte, T. Kratzke, and S. Kaskel, Microporous Mesoporous Mater.
completely hydrated structure. The final structure showed 73, 81 共2004兲.
isotropic contraction of 4.2% in volume, resulting from the 8
C. Prestipino, L. Regli, J. G. Vitillo, F. Bonino, A. Damin, C. Lamberti,
reduction of Cu–O bonds by 1.9%. A. Zecchina, P. L. Solari, K. O. Kongshaug, and S. Bordiga, Chem.
Besides the direct adsorption and separation of gas mol- Mater. 18, 1337 共2006兲.
9
L. Alaerts, E. Seguin, H. Poelman, F. Thibault-Starzyk, P. A. Jacobs, and
ecules, strongly binding species can be used to tune some D. E. De Vos, Chem.-Eur. J. 12, 7353 共2006兲.
properties of MOFs.49 This functionalization approach has 10
S. C. Xiang, W. Zhou, J. M. Gallegos, Y. Liu, and B. L. Chen, J. Am.
been examined in a number of experiments. For example, Chem. Soc. 131, 12415 共2009兲.
11
Bae et al. examined the effect of pyridine-CF3 functionaliza- E. Haldoupis, S. Nair, and D. S. Sholl, J. Am. Chem. Soc. 132, 7528
共2010兲.
tion in Zn2共DMF兲2, where DMF= dimethylformamide.50 12
A. C. McKinlay, B. Xiao, D. S. Wragg, P. S. Wheatley, I. L. Megson, and
They found pyridine-CF3 does not improve the affinity to R. E. Morris, J. Am. Chem. Soc. 130, 10440 共2008兲; F. Bonino, S. Cha-
CO2 compared to DMF functionalized MOF at room tem- van, J. G. Vitillo, E. Groppo, G. Agostini, C. Lamberti, P. D. C. Dietzel,
C. Prestipino, and S. Bordiga, Chem. Mater. 20, 4957 共2008兲.
perature between 0 and 18 bar. This experiment demon- 13
S. Horike, M. Dinca, K. Tamaki, and J. R. Long, J. Am. Chem. Soc. 130,
strated that it is possible to functionalize MOFs with rela- 5854 共2008兲.
14
tively large molecules. The strong binding of pyridine in Q. M. Wang, D. M. Shen, M. Bulow, M. L. Lau, S. G. Deng, F. R. Fitch,
This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP: 147.143.2.5
On: Mon, 22 Dec 2014 04:35:03
094509-12 T. Watanabe and D. S. Sholl J. Chem. Phys. 133, 094509 共2010兲

34
N. O. Lemcoff, and J. Semanscin, Microporous Mesoporous Mater. 55, J. P. Perdew and Y. Wang, Phys. Rev. B 45, 13244 共1992兲.
217 共2002兲. 35
A. Poppl, S. Kunz, D. Himsl, and M. Hartmann, J. Phys. Chem. C 112,
15
B. Xiao, P. S. Wheatley, X. B. Zhao, A. J. Fletcher, S. Fox, A. G. Rossi, 2678 共2008兲; X. X. Zhang, S. S. Y. Chui, and I. D. Williams, J. Appl.
I. L. Megson, S. Bordiga, L. Regli, K. M. Thomas, and R. E. Morris, J. Phys. 87, 6007 共2000兲.
Am. Chem. Soc. 129, 1203 共2007兲. 36
J. Hafner, J. Comput. Chem. 29, 2044 共2008兲.
16 37
C. Lamberti, M. Salvalaggio, S. Bordiga, F. Geobaldo, G. Spoto, A. Y. Wu, A. Kobayashi, G. J. Halder, V. K. Peterson, K. W. Chapman, N.
Zecchina, G. Vlaic, and M. Bellatreccia, J. Phys. IV 7, 905 共1997兲. Lock, P. D. Southon, and C. J. Kepert, Angew. Chem., Int. Ed. 47, 8929
17
K. Hadjiivanov and H. Knozinger, J. Catal. 191, 480 共2000兲. 共2008兲.
18
S. Bordiga, J. G. Vitillo, G. Ricchiardi, L. Regli, D. Cocina, A. Zecchina, 38
J. N. van Niekerk and F. R. L. Schoening, Nature 共London兲 171, 36
Bj. Arstad, M. Bjørgen, J. Hafizovic, and K. P. Lillerud, J. Phys. Chem. B 共1953兲; O. Shekhah, H. Wang, D. Zacher, R. A. Fischer, and C. Woll,
109, 18237 共2005兲. Angew. Chem., Int. Ed. 48, 5038 共2009兲.
19
S. Keskin, J. Liu, R. B. Rankin, J. K. Johnson, and D. S. Sholl, Ind. Eng. 39
G. M. Brown and R. Chidamba, Acta Crystallogr., Sect. B: Struct. Crys-
Chem. Res. 48, 2355 共2009兲. tallogr. Cryst. Chem. B29, 2393 共1973兲.
20
T. Mueller and G. Ceder, J. Phys. Chem. B 109, 17974 共2005兲; F. M. 40
M. J. Frisch, G. W. Trucks, H. B. Schlegel et al., GAUSSIAN 03, Revision
Mulder, T. J. Dingemans, M. Wagemaker, and G. J. Kearley, Chem.
B.05, Gaussian Inc., Pittsburgh, PA, 2003; L. Noodleman, J. Chem. Phys.
Phys. 317, 113 共2005兲; T. Sagara, J. Klassen, and E. Ganz, J. Chem.
74, 5737 共1981兲.
Phys. 121, 12543 共2004兲; S. S. Han, J. L. Mendoza-Cortes, and W. A. 41
D. S. Sholl and J. A. Steckel, Density Functional Theory: A Practical
Goddard, Chem. Soc. Rev. 38, 1460 共2009兲.
21 Introduction 共Wiley, Hoboken, 2009兲; R. M. Martin, Electronic Structure:
T. Sagara, J. Klassen, J. Ortony, and E. Ganz, J. Chem. Phys. 123,
Basic Theory and Practical Methods 共Cambridge University Press, Cam-
014701 共2005兲.
22 bridge, 2004兲.
Y. Gao and X. C. Zeng, J. Phys.: Condens. Matter 19, 386220 共2007兲. 42
23 Q. Ge, S. J. Jenkins, and D. A. King, Chem. Phys. Lett. 327, 125 共2000兲;
T. Sagara, J. Ortony, and E. Ganz, J. Chem. Phys. 123, 214707 共2005兲.
24
D. Dubbeldam, H. Frost, K. S. Walton, and R. Q. Snurr, Fluid Phase S. J. Jenkins, Q. Ge, and D. A. King, Phys. Rev. B 64, 012413 共2001兲.
43
Equilib. 261, 152 共2007兲. S. Keskin and D. S. Sholl, Ind. Eng. Chem. Res. 48, 914 共2009兲.
44
25
N. A. Ramsahye, G. Maurin, S. Bourrelly, P. L. Llewellyn, C. Serre, T. D. Frenkel and B. Smit, Understanding Molecular Simulation: From Al-
Loiseau, T. Devic, and G. Ferey, J. Phys. Chem. C 112, 514 共2008兲. gorithms to Applications, 2nd ed. 共Academic, San Diego, 2002兲; S. Q.
26
M. Nagaoka, Y. Ohta, and H. Hitomi, Coord. Chem. Rev. 251, 2522 Hao, M. Widom, and D. S. Sholl, J. Phys.: Condens. Matter 21, 11
共2007兲. 共2009兲.
45
27
T. Watanabe, S. Keskin, S. Nair, and D. S. Sholl, Phys. Chem. Chem. J. G. Speight, Lange’s Handbook of Chemistry, 16th ed. 共McGraw-Hill,
Phys. 11, 11389 共2009兲. New York, 2005兲.
46
28
X. Cheng, Z. Shi, N. Glass, L. Zhang, J. J. Zhang, D. T. Song, Z. S. Liu, A. Samanta, T. Furuta, and J. Li, J. Chem. Phys. 125, 084714 共2006兲.
47
H. J. Wang, and J. Shen, J. Power Sources 165, 739 共2007兲; J. J. Baschuk F. Birch, Phys. Rev. 71, 809 共1947兲; F. D. Murnaghan, Am. J. Math. 59,
and X. G. Li, Int. J. Energy Res. 27, 1095 共2003兲. 235 共1937兲.
29 48
H. F. Zhu, B. Ka, and F. Murad, World J. Surg. 31, 624 共2007兲; M. R. J. Liu, J. T. Culp, S. Natesakhawat, B. C. Bockrath, B. Zande, S. G.
Miller and I. L. Megson, Br. J. Pharmacol. 151, 305 共2007兲. Sankar, G. Garberoglio, and J. K. Johnson, J. Phys. Chem. C 111, 9305
30
A. J. Kidnay and W. R. Rarrish, Fundamentals of Natural Gas Processing 共2007兲.
49
共CRC, Boca Raton, 2006兲. Z. Q. Wang and S. M. Cohen, Chem. Soc. Rev. 38, 1315 共2009兲; H. X.
31
M. D. Allendorf, R. J. T. Houk, L. Andruszkiewicz, A. A. Talin, J. Pikar- Deng, C. J. Doonan, H. Furukawa, R. B. Ferreira, J. Towne, C. B. Kno-
sky, A. Choudhury, K. A. Gall, and P. J. Hesketh, J. Am. Chem. Soc. bler, B. Wang, and O. M. Yaghi, Science 327, 846 共2010兲.
50
130, 14404 共2008兲. Y. S. Bae, O. K. Farha, J. T. Hupp, and R. Q. Snurr, J. Mater. Chem. 19,
32
K. W. Chapman, G. J. Halder, and P. J. Chupas, J. Am. Chem. Soc. 130, 2131 共2009兲.
51
10524 共2008兲. J. A. Greathouse and M. D. Allendorf, J. Phys. Chem. C 112, 5795
33
G. Kresse and J. Furthmuller, Phys. Rev. B 54, 11169 共1996兲. 共2008兲.

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP: 147.143.2.5
On: Mon, 22 Dec 2014 04:35:03

You might also like