Kinetics and Mechanism of Metal-Organic Framework Thin Film Growth: Systematic Investigation of HKUST-1 Deposition On QCM Electrodes

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Chemical Science Dynamic Article Links < C

Cite this: Chem. Sci., 2012, 3, 1531


www.rsc.org/chemicalscience EDGE ARTICLE
Kinetics and mechanism of metal–organic framework thin film growth:
systematic investigation of HKUST-1 deposition on QCM electrodes†
Vitalie Stavila,*a Joanne Volponi,a Aaron M. Katzenmeyer,a Matthew C. Dixonb and Mark D. Allendorf*a
Received 14th January 2012, Accepted 14th February 2012
DOI: 10.1039/c2sc20065a

We describe a systematic investigation of the factors controlling step-by-step growth of the metal–
organic framework (MOF) [Cu3(btc)2(H2O)3]$xH2O (also known as HKUST-1), using quartz crystal
microbalance (QCM) electrodes as an in situ probe of the reaction kinetics and mechanism. Electrodes
coated with silica, alumina and gold functionalized with OH– and COOH–terminated self-assembled
monolayers (SAMs) were employed to determine the effects of surface properties on nucleation.
Deposition rates were measured using the high sensitivity available from QCM-D (D ¼ dissipation)
techniques to determine rate constants in the early stage of the process. Films were characterized using
grazing incidence XRD, SEM, AFM, profilometry and reflection–absorption IR spectroscopy. The
effects of reaction time, concentration, temperature and substrate on the deposition rates, film
crystallinity and surface morphology were evaluated. The initial growth step, in which the surface is
exposed to copper ions (in the form of an ethanolic solution of copper(II) acetate) is fast and
independent of temperature, after which all subsequent steps are thermally activated over the
temperature range 22–62 " C. Using these data, we propose a kinetic model for the Cu3(btc)2 growth on
surfaces that includes rate constants for the individual steps. The magnitude of the activation energies,
in particular the large entropy decrease, suggests an associative reaction with a tight transition state.
The measured activation energies for the step-by-step MOF growth are an order of magnitude lower
than the value previously reported for bulk Cu3(btc)2 crystals. Finally, the results of this investigation
demonstrate that the QCM method is a powerful tool for quantitative, in situ monitoring of MOF
growth in real time.

Introduction technologies for implementing MOFs in sensors and other


devices is the assembly of dense films or coatings on various
Metal–organic frameworks (MOFs) have attracted considerable substrates.10 While synthesis of MOFs as bulk powders is well
interest as a result of their tunable pore structures and chemical developed, there is a lack of reliable protocols for growing MOFs
functionalities.1–5 Their high pore volume and surface area create on surfaces (also called SURMOFs, surface-mounted MOFs12,13)
opportunities for use in gas storage,6 catalysis,7 and small- of a desired thickness and morphology, while maintaining their
molecule separation.8 The ability of MOFs to reversibly absorb functionality. Understanding how MOF thin films form is
small molecules also makes them attractive for sensing applica- therefore a crucial aspect of expanding the repertoire of MOFs
tions, a topic that was recently reviewed.9 Their highly ordered that can be deposited and for developing reproducible and
structure also suggests possibilities for fabricating electronic practical deposition methods.
devices with nanoscale features.10 In these latter two applica- Recently, we demonstrated the concept of stress-induced
tions, MOFs have advantages relative to organic polymers and chemical sensing of small molecules using thin films of the MOF
other nanoporous materials due to their relatively high thermal [Cu3(btc)2(H2O)3]$xH2O (referred to henceforth as Cu3(btc)2,
stability, uniform pore structure, and potential to be tailored to but also known as HKUST-1;14 btc ¼ 1,3,5-benzenetricarbox-
enhance uptake of specific analytes or materials such as catalytic ylate) deposited on microcantilevers.15 Previous work showed
or luminescent nanoparticles.11 One of the most critical enabling that Cu3(btc)2 can efficiently absorb a wide variety of molecules,
including H2O,16 NH3,17 CO2,18 CH4,19 C2H2,20 to name just
a
a few. Cu3(btc)2 also displays a capacity equal to or greater than
Sandia National Laboratories, Livermore, CA, 94551, USA. E-mail:
activated carbons in removing such harmful gases as sulfur
vnstavi@sandia.gov; mdallen@sandia.gov
b
Biolin Scientific, Inc., Linthicum Heights, MD, 21090, USA dioxide, tetrahydrothiophene, benzene, dichloromethane,
† Electronic supplementary information (ESI) available: General ethylene oxide and carbon monoxide.21 These properties suggest
experimental aspects, details of QCM measurements, and additional that this MOF could be useful in certain sensing applications,
characterization of Cu3(btc)2 coatings. See DOI: 10.1039/c2sc20065a

This journal is ª The Royal Society of Chemistry 2012 Chem. Sci., 2012, 3, 1531–1540 | 1531
and indeed, we recently demonstrated that coatings of Cu3(btc)2 Herein we describe a systematic investigation of the role of
on surface acoustic wave (SAW) sensors can be used to detect deposition conditions (temperature, concentration and
water vapor at ppm levels.22 Our sensing devices were coated substrate) on Cu3(btc)2 growth kinetics and film properties, with
using the layer-by-layer (LBL) method developed by W€ oll and the objective of gaining new insight into the elementary reactions
co-workers,23,24 a versatile technique that has been used to create leading to film growth, that is relevant to the development of
films of several MOFs.3 Although a number of reports describe MOF-based coating processes. Reaction dynamics were
the growth of Cu3(btc)2 on surfaces, most of these are focused on obtained from QCM high-resolution frequency measurements,
new deposition methods and not on the growth kinetics. For which are a direct indicator of the rate of MOF film growth, as
example, thin-film growth of Cu3(btc)2 on various substrates, described above. The structure and morphology of the resulting
including Al2O3,25 SiO226 and various self-assembled monolayers MOF films were probed via grazing incidence X-ray diffraction
(SAMs),13,23,24,27–33 has been demonstrated. A growth mechanism (GIXRD), reflection–absorption infrared spectroscopy (RAIR),
was proposed by Shekhah et al.,13 and it is believed to involve and surface and cross-sectional imaging (SEM, white light
stepwise ligand exchange reactions between pre-formed interferometry, and AFM). A central feature of this study is the
secondary building units (SBUs) and anchored surface correlation of information from these different techniques to
groups.2,23,24,30,34–36 However, none of these studies report actual generate a coherent model for MOF film composition, texture,
growth rates and do not address the influence of processing and dynamics at various interfaces. Findings pertinent to the
conditions on the growth rate and film properties. growth of Cu3(btc)2 on carboxyl- or hydroxy-terminated Au/
Quantification of growth rates and the establishment of alkanethiol SAMs, SiO2 and Al2O3 coated electrodes are pre-
a kinetic mechanism require an accurate time-resolved method to sented. We compare the growth of Cu3(btc)2 in different envi-
monitor the growth process. Shekhah et al. used surface plasmon ronments (solution, substrate, temperature) and show that there
resonance (SPR) to monitor the deposition of Cu3(btc)2 on Au are substantial differences between the mechanism of growth on
surfaces functionalized with carboxylate-terminated SAMs. surfaces vs. the processes controlling bulk powder synthesis.
They found that sequential addition of Cu2(OAc)4 and H3btc Moreover, we are able to distinguish between an initial nucle-
(H3btc ¼ 1,3,5-benzenetricarboxylic acid) leads to a stepwise ation step and subsequent ‘‘steady-state’’ reactions. The resulting
mass increase, yielding highly homogenous and crystalline data fill an important gap in the understanding of this process,
Cu3(btc)2 coatings.13,24 However, SPR, being an optical tech- which ultimately will be useful in developing larger-scale pro-
nique, is best suited for the measurement of relatively thin cessing methods for integrating Cu3(btc)2 with other materials.
coatings; in addition, to provide quantitative data about the
thickness and mass of the adsorbed layer, the refractive index of
the MOF film must be known. In the investigation described Results
here, we employed a quartz crystal microbalance (QCM), i.e. an
Growth rates
acoustic technique, to monitor the deposition process. This
method does not require knowledge of the refractive index and The LBL growth method described by W€ oll, Fischer and co-
allows measurement of thicker films (up to several microns), workers is a step-by-step growth method23,24 that in the case of
which may be necessary for membrane or sensing applications. Cu3(btc)2 consists of sequentially flowing ethanolic solutions of
The QCM technique is based on the piezoelectric property of the metal precursor (Cu2(OAc)4) and organic linker (H3btc) over
quartz, in which a mechanical shear oscillation can be induced by a substrate with solvent washing steps between cycles. During the
an alternating electric field. As molecules absorb onto the elec- MOF deposition process ligand exchange reactions take place at
trodes (Au or other metals), the oscillation frequency of the the interface between the solid and liquid phases, allowing the
quartz crystal decreases, allowing in situ monitoring of the metal ions to bind to linker groups at the surface and vice
deposition kinetics. The resolution of QCM measurements is on versa.13,23,24,30 Although the QCM is sensitive to extremely small
the order of #1.0 ng cm$2, which is the reason the technique is mass changes, realizing the maximum sensitivity of the technique
also termed ‘‘quartz crystal nanobalance’’. The upper limit of film can be difficult, especially with liquids. Typical environmental
thickness that the method can be used to measure is several effects associated with QCM measurements in solutions include:
microns, depending on the viscoelastic properties of the applied (i) temperature changes during the experiment; (ii) quality of the
material.37 For rigid films, the Sauerbrey eqn (1) is often used to surface (deposition on the QCM electrode can be affected by the
correlate the observed frequency shifts to deposited mass:38 roughness, porosity, hydrophilicity, etc.); (iii) complications can
pffiffiffiffiffiffiffiffiffiffi be introduced when viscoelastic films are deposited; (iv) adhesion
A rq Gq
Dm ¼ Df (1) of the film to the surface may affect the measurement accuracy,
$2f02
especially when accumulated stress is present causing films to
where Dm is the change in mass, A is the surface area of the peel off or partially lift off. In our experiments, these sources of
resonator, rq and Gq are the density and shear modulus of error were minimized by strict temperature control, paying
quartz, f0 is the resonance frequency of the unloaded resonator careful attention to the QCM crystal surface functionalization,
and Df is the change in resonance frequency. Although several growing relatively thin coatings of a rigid MOF films (vide infra),
authors describe the use of the QCM technique to characterize and optimizing the flow rates used during deposition to minimize
the absorption/desorption characteristics of various molecules instrument noise (a schematic representation of the experimental
on MOFs,32,39–41 no systematic investigations using this tech- set-up is presented in Fig. S1, ESI†).
nique to monitor the deposition of MOF thin films42 have Of the three surfaces examined in our investigation, silica-
appeared. coated electrodes are the ones most relevant for a detailed kinetic

1532 | Chem. Sci., 2012, 3, 1531–1540 This journal is ª The Royal Society of Chemistry 2012
investigation because there is no intervening SAM to complicate QCM-D system operated in pulsed mode with in situ dissipation
the interface chemistry (growth on SAMs was also previously monitoring and Peltier temperature control was used. Measuring
studied in detail13,23,24,30,36,43,44). Moreover, for sensing purposes, the oscillatory decay when the electric field applied to the QCM
growing Cu3(btc)2 on this material is attractive because the crystal is turned off provides information concerning the amount
strong covalent binding to surface hydroxyl groups is much more of energy dissipated by the MOF layer. The dissipation (D) is
stable thermally than thiol-based SAMs. Finally, compared with inversely proportional to the decay time, s (eqn (2)):
alumina, which requires an additional processing tool (either 1
atomic-layer deposition or reactive sputtering), growth of silicon D¼ (2)
pf s
oxide is a standard tool available in microelectronics fabrication
facilities, making it the most practical interface material for where f is the crystal oscillating frequency. For a soft film, the
sensing purposes. We therefore confine our mechanistic analysis decay time is small due to increased coupling with the
to growth on silica; however, as will be seen below, the overall surrounding medium, leading to higher dissipation, whereas for
reaction kinetics following the initial interface step are very a rigid film, the decay time is large leading to smaller dissipation.
similar among the three surfaces, indicating that once growth is After five cycles of alternatively flowing ethanolic solutions of
initiated, most likely by the formation of the first SBU, it Cu(OAc)2 and H3btc over a silica-coated QCM-D crystal the
proceeds in the same way regardless of the underlying substrate. total frequency change is 87.5 Hz, while the dissipation remains
All three deposition surfaces are amorphous, which is again close to zero (Fig. 1). The relatively low dissipation suggests that
typical of the morphology of these materials available from the MOF film is rigidly attached to the electrode and there is
standard microelectronic processing methods. almost no viscoelastic contribution to the frequency change; thus
The frequency change detected during the step-by-step depo- the Sauerbrey equation is valid for this material.
sition of Cu3(btc)2 on a silica-coated QCM crystals is shown in At the beginning of a deposition experiment pure EtOH is
Fig. 1, which illustrates the high quality of data possible using the circulated over the QCM-D crystal, resulting in a relatively flat
QCM-D instrument. Although the flow rate is constant during baseline. Growth is initiated by the introduction of ethanolic
the experiment, changing from one reactant to another yields Cu(OAc)2, which causes a sharp drop in frequency that reaches
a measurable and distinct frequency change. The overall decrease a plateau in #2 min (Fig. 1). As suggested previously, this
in resonant frequency correlates with the added mass of behavior indicates that a finite number of adsorption sites is
adsorbing species; no detectable change in frequency is observed available.45 The difference in frequency between the EtOH
when pure ethanol is introduced into the system. The stepwise baseline and the plateau is about 4 Hz, which translates into an
changes in frequency correspond to alternate solutions being added mass of #72 ng cm$2 using eqn (1). After the frequency
circulated over the QCM-D crystal and indicate a mass increase change levels off, pure EtOH is passed over the QCM electrode,
over time. A relatively simple relationship between the frequency resulting in a slight increase in frequency, presumably due to
change and mass uptake is given by the Sauerbrey equation (eqn some dissociation of loosely adsorbed copper(II) species. Based
(1)). However, since there are cases when the Sauerbrey equation on the rapid frequency change, it appears that the adsorption of
does not hold (e.g. when the added mass is not rigidly and evenly solution-phase paddle-wheel Cu2(OAc)4 units at the solid–liquid
deposited on the electrode surface), we assessed the impact of the interface is fast. Copper(II) is known to have high affinity to
MOF viscoelestic properties by measuring the energy dissipation various surfaces and the overall dynamics of copper(II) salts
of the QCM crystals during film growth (Fig. 1). A Q-Sense! depositing onto surfaces may be diffusion- or adsorption-rate
controlled.46 Although the QCM measurements do not directly
identify the species deposited, we can determine this from the
mass gain/cycle and the density of reactive sites per unit area
(obtained from the crystal structure). Shekhah et al. previously
suggested that growth of Cu3(btc)2 on OH-terminated SAMs
occurs preferentially along the (111) crystallographic direction
and involves paddle-wheel SBUs. Using the same logic and
assuming that the observed mass increase during the copper(II)
step is due to the adsorption of intact Cu2(OAc)4 units, this
results in 0.181 nmol Cu2(OAc)4 per cm2 or 1.1 % 1014 molecules
cm$2. Even if Cu2(OAc)4 units require more than one surface
active site (e.g. OH groups) to bind, the number of available
hydroxyl groups for similarly treated SiO2 surfaces was reported
to be as high as 5 OH groups nm$2 or 5 % 1014 surface sites
cm$2.47 For a complete monolayer surface coverage assuming
that the subsequent MOF growth occurs along the (111) crys-
tallographic direction, the expected mass gain was calculated to
be 76 ng cm$2, which is close to the measured value of 72 ng cm$2
Fig. 1 QCM-D results showing changes for various harmonics of the for the first copper step. This is an indication that the initial
frequency and dissipation after five Cu(OAc)2–H3btc deposition cycles growth involves paddle-wheel Cu2(OAc)4 species.
on SiO2-coated electrodes at 15 " C. The flow rate was maintained Following the initial exposure to Cu2(OAc)4, the stepwise
constant at 100 mL min$1 throughout the experiment. process continues by passing an ethanolic solution of trimesic

This journal is ª The Royal Society of Chemistry 2012 Chem. Sci., 2012, 3, 1531–1540 | 1533
acid over the electrodes. This is again accompanied by a signifi-
cant decrease in frequency. However, the rate of this frequency
change is slower than that of the copper(II) step. Conceptually,
incorporation of the btc linker into the film probably involves
a metathesis reaction between the coordinated acetate groups on
the surface and solvated trimesic acid, which we expect will
involve a specific transition state and thus be thermally activated.
For the second, third, fourth and fifth Cu(II) steps the measured
mass changes (86, 96, 99 and 107 ng cm$2, respectively) are
slightly larger compared to the first step. The calculated mass
changes during the first five H3btc steps are 255, 230, 224, 214
and 207 ng cm$2, respectively. This decrease in H3btc : Cu molar
ratio as the cycle progress may be due to differing ratios on the
surface between the two precursors during the nucleation and
film growth stages (vide infra). For each of the H3btc steps the
frequency change and the corresponding Sauerbrey mass change Fig. 2 Sauerbrey mass change after three Cu(OAc)2–H3btc deposition
is larger than the amount strictly required by the stoichiometry of cycles on SiO2-coated electrodes at 15, 30 and 45 " C. The inset shows the
the reaction of Cu2(OAc)4 with H3btc. This suggests that during corresponding Eyring–Polanyi plots for each of the individual deposition
the H3btc step more linker molecules are adsorbed on the surface steps.
than required by the 3 : 2 Cu : btc stoichiometry of HKUST-1,
or alternatively, solvent (EtOH), linker, and other molecules
temperatures studied, Cu(OAc)2 steps subsequent to the first one
(H2O, AcOH) are trapped inside the newly formed MOF pores,
produce slightly larger frequency changes. Higher temperatures
contributing to the observed QCM frequency change. The
also give larger frequency changes, thus higher mass uptake.
increase in mass change upon each subsequent cycle may also be
After three Cu(OAc)2/H3btc cycles the total mass added is 0.97,
explained by an increase in surface coverage or film surface area.
1.18 and 1.49 mg cm$2 at 15, 30 and 45 " C, respectively. Since the
Although no structural information for the five cycle-coated
concentration of the reagents during the flow cycle is essentially
QCM crystal is available, the presence of copper on the surface
unchanged (only a small fraction of the total Cu(II) or linker
was confirmed by energy dispersive spectroscopy (EDS) (Fig. S2,
available in the solution is consumed by the surface reactions),
ESI†).
one can assume a pseudo-first order reaction for the MOF
The relationship between the deposited film thickness (Dh) and
growth process, which is only dependent on the concentration of
Sauerbrey mass is given by the following relationship (eqn (3)):
reactive sites on the surface. This is consistent with the QCM-D
1 measurements showing that, after an initial fast mass uptake, the
Dh ¼ Dm (3)
d mass uptake saturates.
in which Dm is the mass per cm2 obtained from eqn (1) and d is
the density of film. As mentioned above, after five combined
Cu(OAc)2/H3btc cycles the total frequency change is 87.5 Hz, Kinetic analysis
which corresponds to #1.34 mg cm$2 added mass. Assuming the
film has the density of bulk hydrated Cu3(btc)2 (0.96 g cm$3),14 The growth rate corresponding to the individual cycles in Fig. 2
the deposited film thickness calculated from eqn (3) is about was determined from the QCM-D frequency changes (Table 1)
13.9 nm, or an average of 2.8 nm per step. and plotted according to the Eyring–Polanyi equation, which is
Measurements of the effect of temperature on the Sauerbrey similar to the empiric Arrhenius equation, but it has a theoretical
mass after the first three Cu(OAc)2–H3btc cycles allow us to basis as it follows directly from transition-state theory, which can
distinguish between a rapid nucleation process and a slower two- be extended to condensed-phase reactions:48
step growth reaction that occurs during subsequent cycles. As " #
kB T DG‡
seen in Fig. 2, in all reaction steps except the first exposure to k¼ exp $ (4)
h RT
Cu(OAc)2 solution, the overall growth rate increases as the
temperature is increased from 15 to 45 " C. However, the mass where k is the reaction rate constant at temperature T, DG‡ is the
uptake during the initial copper(II) exposure is almost indepen- Gibbs energy of activation, kB is Boltzmann’s constant and h is
dent of temperature. Combining this observation with the shape Plank’s constant. In contrast to the Arrhenius equation, which is
of the mass uptake curve (a sharp rise followed by a plateau) based on the empirical observation that the rate of a chemical
suggests that copper initially deposits on silica surfaces via a fast reaction increases with temperature, the Eyring equation is
(zero activation energy) sticking reaction that is limited by the a theoretical construct based on transition state theory.45 Since
availability of reactive surface sites. If so, the dependence on the DG‡ ¼ DH‡ $ TDS‡, eqn (4) can be represented as
Cu(II) species should be zeroth order, a fact that was confirmed " # " ‡#
kB T DH ‡ DS
experimentally. k¼ exp $ exp (5)
h RT R
Following the fast copper initiation step, the btc linker
undergoes a slower, thermally activated reaction in the second where DH‡ and DS‡ are the enthalpy and entropy of the transi-
half of the cycle (Fig. 2). Interestingly, at each of the three tion state. The linear form of eqn (5) is:

1534 | Chem. Sci., 2012, 3, 1531–1540 This journal is ª The Royal Society of Chemistry 2012
Table 1 Reaction rates, calculated activation enthalpy (DH‡), entropy (DS‡), pre-exponential factors (A) and activation energy (Ea) for the Cu3(btc)2
step-by-step growth

Reaction rate/s$1

Reaction coordinate 15 " C 30 " C 45 " C DH‡/kJ mol$1 DS‡/J K$1 mol Ea/kJ mol$1 A/s$1

Cu(OAc)2, step 1 2.01 % 10$8 2.02 % 10$8 2.08 % 10$8 — — — —


Cu(OAc)2, step 2 2.49 % 10$8 3.34 % 10$8 4.69 % 10$8 13.5 $343.3 16.0 ' 0.5 1.99 % 10$5
Cu(OAc)2, step 3 2.58 % 10$8 3.42 % 10$8 4.61 % 10$8 12.2 $347.6 14.7 ' 0.5 1.19 % 10$5
H3btc, step 1 1.03 % 10$8 1.42 % 10$8 2.14 % 10$8 16.0 $342.2 18.5 ' 0.6 2.30 % 10$5
H3btc, step 2 1.05 % 10$8 1.47 % 10$8 2.31 % 10$8 17.4 $337.0 19.9 ' 0.6 4.25 % 10$5
H3btc, step 3 1.01 % 10$8 1.60 % 10$8 2.38 % 10$8 19.2 $330.8 21.7 ' 0.6 8.98 % 10$5

k DH ‡ 1 kB DS‡ 32 " C, with constant solution flow rates of 0.1 mL min$1. Fig. 3
ln ¼$ þ ln þ (6) shows the mass change after 20 cycles of QCM crystal exposure
T R T h R
to ethanolic solutions of Cu2(OAc)4 and H3btc. For all the
We applied the Eyring–Polanyi formalism described above to substrates investigated the Cu3(btc)2 growth is essentially linear,
determine the enthalpy, entropy, pre-exponential factor and although the noise level is somewhat higher in this instrument.
activation energy for the first three Cu(OAc)2–H3btc cycles The fastest growth rate is observed for alumina-coated crystals,
(Table 1). The initial Cu(II) deposition step (step 1) has a rate that while the Cu3(btc)2@COOH-terminated SAMs display the
is independent of the Cu(OAc)2 concentration and temperature, slowest growth rate of the four substrates.
which is typical of adsorption reactions in which no rearrange- The identity of the Cu3(btc)2 films deposited on the four
ment or bond breaking in the adsorbate is involved. In contrast, substrates was confirmed by ex situ XRD and RAIR spectros-
the next steps in the coating process have a distinct temperature copy measurements (Fig. S3, ESI†). Fig. 4 compares the XRD
dependence. Assuming a pseudo-first-order reaction, we calcu- patterns for Cu3(btc)2 coatings obtained on COOH, OH, SiO2
lated the deposition rates for each of the Cu(II) and H3btc indi- and Al2O3 surfaces with the indexed diffraction pattern of bulk
vidual coating steps. The resulting activation energies for the HKUST-1. The collected GIXRD patterns show characteristic
second and third Cu(II) deposition steps were calculated to be features of crystalline Cu3(btc)2, with some broadening of the
16.0 and 14.7 kJ mol$1, respectively. Slightly larger activation peaks compared to bulk material. The growth on COOH-
energies are observed for the three H3btc deposition steps: 18.5, terminated SAMs results in a film with a preferred orientation
19.9 and 21.7 kJ mol$1. These low Ea values are consistent with along the (100) direction. In contrast, a significant preferred
concerted (and likely solvent-assisted) reactions in which bond orientation along the (111) crystallographic direction is observed
breaking and formation occur simultaneously. for Cu3(btc)2 films grown on OH-terminated SAMs. This agrees
The entropy of activation DS‡ is determined from the intercept with W€ oll and co-workers, who first reported that the terminal
of the ln(k/T) vs. 1/T plot at y ¼ 0. The calculated values are functional groups can govern the Cu3(btc)2 film growth.13,31
between $343.3 to $347.6 J mol$1 K for the Cu(II) steps and Thus, thin films of Cu3(btc)2 grown using the step-by-step
between $330.8 and $342.2 J mol$1 K for the H3btc steps procedure on COOH and OH-terminated SAMs are highly
(Table 1). These large negative values lead to small pre-expo- oriented along the (100) and (111) direction, respectively. The
nential factors A. Generally, an entropy decrease is expected as
a result of the loss of translational and rotational degrees of
freedom when the reactant leaves the liquid phase and binds to
the surface. However, they also suggest a ‘‘tight’’ transition state,
in which the reactants must adopt a specific geometric orienta-
tion for reaction to occur. This is consistent with the metathesis-
type reaction that would occur during acetate/btc and btc/acetate
exchange (vide infra).

Substrate effects and film morphology


The effects of the substrate on the ‘‘global’’ activation energy
corresponding to multiple cycles of step-by-step Cu3(btc)2
growth were investigated using a SRS-200 QCM system, which
monitors only changes in fundamental frequency. Since the
QCM-D measurements reveal that the added MOF is tightly
bound to the surface of the crystals under the experimental
conditions tested, the dissipation component is insignificant and
can be omitted. MOF growth on four types of substrates was
explored: Al2O3, SiO2, and Au functionalized with COOH- and
OH-terminated SAMs. In order to compare the deposition rates Fig. 3 Cu3(btc)2 deposition on Al2O3, SiO2, and Au functionalized with
of Cu3(btc)2 on various substrates the growth was performed at COOH and OH SAMs at 32 " C.

This journal is ª The Royal Society of Chemistry 2012 Chem. Sci., 2012, 3, 1531–1540 | 1535
difference might originate in differing surface morphologies for
the two substrates, which potentially could have various
concentrations of active sites on the surface, resulting in
a coverage dependence for the adsorption of reactive precursors
from the solution. If this is the case, some difference in the
surface properties of the coatings might be observable. Addi-
tional characterization was performed to test this hypothesis.
First, a comparative analysis of the GIXRD results reveals that
the MOF film on silica has a slightly larger degree of texture
along the (100) crystallographic direction, which may indicate
a more homogeneous growth. Second, the surface roughness of
the Cu3(btc)2@Al2O3 film is greater than that of the film grown
on silica. The measured RMS surface roughness of
Cu3(btc)2@SiO2 film is 6.4 nm, while the corresponding value for
Cu3(btc)2@Al2O3 film is 10.0 nm (Fig. S4, ESI†). This is
consistent with an overall higher rate of growth on alumina.
Additional evidence supporting this explanation is that the
Fig. 4 Grazing incidence XRD diffraction patterns at u ¼ 0.2" for growth rate on alumina increases with time, particularly at 52
Cu3(btc)2 films deposited on COOH-terminated SAM, OH-terminated and 62 " C, resulting in noticeable curvature in the mass vs. time
SAM, SiO2 and Al2O3 surfaces, in comparison with bulk HKUST-1. For plot (Fig. 5b). The deviation from the linearity at higher
clarity, only selected hkl reflections of bulk Cu3(btc)2 are labeled. temperatures does not necessarily mean that a different growth
mechanism is operative; instead, it may arise from the deposition
of Cu3(btc)2 on non-planar surfaces. Rougher films presumably
reason for the different film texture lies in the enhanced inter- have higher surface areas; as growth proceeds, the surface area
actions of the COOH surface groups with the paddle-wheel Cu2 may actually increase, leading to an increase in the rate with time,
dimeric units in the (100) plane, while the OH groups exhibit similar to the observed growth of self-propagating molecular
preferential binding to the apical position of the Cu2+ ions situ- assemblies.49 Unfortunately, since we cannot control the surface
ated in the (111) lattice plane.24,31 In contrast with these previous area or reactive site density of our substrates, this explanation is
results,13,31 Cu3(btc)2 growth on QCM crystals using continuous- at best speculative at this point.
flow conditions results in less oriented films with some preferred As mentioned above, the smoothest HKUST-1 films were
orientation along the (100) direction for COOH-terminated grown on silica surfaces. We also observe that the full-width at
SAMs and (111) texture for OH-terminated SAMs. Shekhah half-maximum (FWHM) of the Cu3(btc)2@SiO2 film diffraction
et al. observed a similar behavior by correlating the differing film peaks is narrower than that of films grown on other substrates,
orientations obtained for Cu3(btc)2 grown on carboxylate- vs. which may indicate to higher film crystallinity. It is therefore
hydroxyl-terminated SAMs with the composition of the growth tempting to correlate surface roughness with the crystallinity (or
surface expected from the crystal structure.13 As far as even the extent of disorder) of the films. Unfortunately, this is
Cu3(btc)2@SiO2 and Cu3(btc)2@Al2O3 coatings are concerned, difficult to do reliably, since the film roughness determined by
in both cases the films are polycrystalline, with a slight preferred AFM is due to small nanometre-sized surface features, while the
orientation along the (111) direction. Oxygen plasma treatment total film thickness contributing to the observed diffraction
of SiO2 and Al2O3-coated QCM crystals is known to produce pattern is much larger, of the order of 100 nm. Furthermore, we
highly hydroxylated {Si–OH} and {Al–OH} surfaces.46 These note that the observed FWHM of a sample’s GIXRD can be
groups have remarkably high polarity and most likely interact influenced by a number of factors, such as grain size and micro
with the water molecules of the [Cu2(CH3CO2)4(H2O)2] paddle- strain of the film. Thus, a much more detailed experimental effort
wheel secondary building units (SBUs), facilitating the growth will be required to determine what relationship, if any, exists
along the (111) crystallographic direction. between bulk film crystallinity and surface order, which is
The difference in global deposition rates on silica and alumina outside the scope of this initial investigation of HKUST-1
extends to the temperature dependence, which was determined growth kinetics.
from Eyring–Polanyi plots obtained over a large number of A previous study by Khan et al. reported an activation energy
growth cycles and five temperatures (20–62 " C; Fig. 5a and b). As value of 133 kJ mol$1 for bulk Cu3(btc)2 growth from Cu(NO3)2
in the case of the QCM-D measurements, higher temperature under solvothermal conditions.50 This value is consistent with the
accelerated the MOF growth for both SiO2 and Al2O3 substrates. observations of Shekhah et al.,13 who found that HKUST-1
The overall activation energy for Cu3(btc)2 film growth was growth from the nitrate salt is much slower than from the acetate,
derived from the Eyring–Polanyi plots in Fig. 5 and is 14.6 ' presumably because pre-formed SBUs do not exist in solution.
1.2 kJ mol$1 for silica and 11.9 ' 1.1 kJ mol$1 for alumina. This The global activation energies of 14.6 ' 1.2 and 11.9 ' 1.1 kJ
difference is consistent with the larger mass uptake during mol$1 that we observe for Cu3(btc)2 thin film growth on silica and
Cu3(btc)2 growth on alumina compared to silica. alumina, respectively, support this hypothesis, being about an
The reason for this unexpected, but statistically significant, order of magnitude lower than the bulk value. This suggests that
difference in activation energy between the two oxides under growth of HKUST-1 on surfaces from nitrate might occur at
similar flow rates and reactant concentrations is unclear. This reasonable rates under solvothermal conditions. Unfortunately,

1536 | Chem. Sci., 2012, 3, 1531–1540 This journal is ª The Royal Society of Chemistry 2012
Fig. 5 Temperature-dependent studies of Cu3(btc)2 deposition of on silica-coated (a) and alumina-coated (b) QCM crystals. The activation energies
extracted from the Eyring–Polanyi plots are 14.6 and 11.9 kJ mol$1 for SiO2- and Al2O3-coated electrodes.

at these temperatures SAMs on Au would decompose. More extracted from QCM measurements using eqn (1) and (3) are
importantly, at such high temperatures, it is very difficult to about 75% of the SEM values, indicating that the actual film
confine growth to the surface and control the film thickness. density is lower than the single-crystal Cu3(btc)2 density assumed
Typical SEM images of the Cu3(btc)2 coatings on various in the analysis of the QCM data. Indeed, filling the Cu3(btc)2
substrates after 40 cycles of step-by-step deposition are shown in pores with EtOH will likely results in a lower density film,
Fig. 6a–d. Fairly uniform films are observed, free of cracks and compared to a film filled with water molecules. In addition, the
fissures, indicating that the MOF film is likely dense. Fig. 6e as-synthesized MOF film may have some additional porosity and
shows the transversal cross-section micrograph of the 40-cycle empty space, which will further reduce the actual density of
Cu3(btc)2 coating on the alumina, indicating an average MOF the film.
film thickness of about 100 nm. The film thickness values A detailed examination of the MOF surface by AFM and
profilometry reveals slight differences in surface morphology of
the films prepared on different substrates. The AFM images of
the Cu3(btc)2 films deposited on silica have a regular smooth
surface and dense structure (Fig. 7). The smoothest films were
obtained on silica, with an average surface roughness of about
6.4 nm, which corresponds to a step height of approximately two
to three unit cells.14 Surface interferometry scans over a larger
area of the QCM crystals (tens of microns) for Cu3(btc)2@SiO2
also indicate relatively smooth surfaces (Fig. S5, ESI†). Higher
surface roughness is observed in the case of Cu3(btc)2@Al2O3
and Cu3(btc)2 deposited on both COOH and OH-terminated
SAMs (between 16 and 27 nm). The initial surface of the silica-
coated electrodes was also the smoothest, however, with RMS
roughness of less than 1 nm. The Au and alumina surfaces had
slightly higher roughness, between 2 and 3 nm. Step-height AFM
measurements (Fig. 7e) on Cu3(btc)2@SiO2 surfaces indicate
a step height between 2 and 3 nm, which correlates well with the
average film thickness extracted from QCM-D measurements.
However, this value is larger than what is expected for the
deposition of a single pre-formed SBU (1.1 to 1.5 nm, depending
on the direction of growth).51–54 This discrepancy may be due to
the fact that more than one SBU per cycle is deposited; more-
over, the deposited films are highly hydrated and possibly
contain solvent molecules at the surface interphase.

Deposition mechanism
The results of this investigation, combined with previous studies
of HKUST-1 growth in the literature,13,23,24,42,55 indicate that the
Fig. 6 SEM images of Cu3(btc)2 coatings (40 cycles) on SiO2 (a), Al2O3 step-by-step deposition of Cu3(btc)2 on surfaces involves the
(b), COOH-terminated Au SAM (c), OH-terminated SAM (d) and the following steps: (i) nucleation, in which Cu(II) species (most likely
transversal cross-sectional SEM image of the Cu3(btc)2@Al2O3 coating (e). as paddle-wheel Cu2(OAc)4 units) rapidly adsorb on the surface

This journal is ª The Royal Society of Chemistry 2012 Chem. Sci., 2012, 3, 1531–1540 | 1537
coordinated OAc$ ligands by surface OH, this reaction is likely
quite exothermic and perhaps solvent assisted as well. The very
early stages of adsorption probably involve conformationally
disordered Cu2(OAc)4 units randomly oriented on the surface.
As the surface becomes saturated with Cu(II) species, it is possible
they adopt a more uniform packing and become more oriented.
Subsequent exposure to a solution of the organic linker results in
ligand exchange at rate k1 and possible further reorientation of
the surface Cu2(OAc)4 to accommodate the bridging btc3$
ligands. The QCM measurements indicate that at the initial stage
of this reaction, the btc : Cu molar ratio is higher than in
subsequent cycles. This could be rationalized assuming that all
four bridging acetate ligands in the surface-bound Cu2(OAc)4
units are replaced by btc3$ groups during the initial H3btc steps,
resulting in a 2 : 1 btc : Cu molar ratio. Consequently, it is during
the next exposure to Cu2(OAc)4 (occurring at rate k2) that the
bridging btc3$ ligands are replaced to create SBUs with the 2 : 3
btc : Cu molar ratio of the MOF structure. Finally, as the growth
proceeds, sequential exposure to solutions of copper(II) acetate
and trimesic acid results in Cu3(btc)2 nuclei that coalesce to
generate a continuous film. Although growth on individual
crystal surfaces is epitaxial during growth on the substrates used
here, the resulting films are polycrystalline and have differing
degrees of texture along the (111) or (100) crystallographic
directions. This is in contrast with previous reports of Cu3(btc)2
growth on SAM-functionalized surfaces, which were highly
oriented.13,31 However, in those cases, the initial substrate was
Fig. 7 AFM images of Cu3(btc)2 films deposited on SiO2 (a), Al2O3 (b), single-crystal silicon, which is highly oriented and much
COOH-functionalized SAM (c), OH-functionalized SAM (d) and the smoother than the surfaces used here.
step height of the Cu3(btc)2@SiO2 coating after 1 cycle (e). Although it is not feasible to determine the detailed geometry
of the reaction pathway and transition states with any precision
from our measurements, it is evident that in both the Cu(II) and
(k0); (ii) ligand exchange between coordinated acetate and tri- btc addition steps an exchange of acetate and btc anions must
mesic acid (k1); (iii) Cu2(OAc)4 addition (k2), resulting in occur. Furthermore, the mass uptake determined from the QCM
Cu3(btc)2 structure formation; and (iv) continuous film growth frequency changes during the copper(II) steps does not suggest
via k1 and k2, with Cu3(btc)2 crystallite coalescence to form dissociation of the paddle-wheel Cu2(OAc)4 units into mono-
a dense film. Fig. 8 is a schematic representation of the proposed mers. Rather, the similar transition state thermodynamics
mechanism of Cu3(btc)2 film nucleation and growth, using the (Table 1) suggest a high degree of similarity between the Cu(II)
example of a hydroxylated oxide surface (e.g. silica). During the and btc addition reactions. Shekhah et al. reached a similar
nucleation step (rate k0), surface OH groups react with conclusion by correlating the differing film orientations obtained
Cu2(OAc)4 species in solution, creating the initial surface for Cu3(btc)2 grown on carboxylate- vs. hydroxyl-terminated
precursor to MOF growth. Our results suggest that this process SAMs with the composition of the growth surface expected from
begins immediately after exposure to the Cu(II) solution and the crystal structure.13 Reasoning from the results in Table 1, at
surface saturation is reached within minutes. Since this process H3btc/step 1, the linker likely encounters Cu(II) ions that are
has no activation energy, but apparently involves hydrolysis of coordinated to acetate groups following the initial exposure to

Fig. 8 Schematic representation of the proposed model for Cu3(btc)2 nucleation and growth on oxide surfaces. The atoms are shown as follows: Cu –
green, O – red, C – gray.

1538 | Chem. Sci., 2012, 3, 1531–1540 This journal is ª The Royal Society of Chemistry 2012
the Cu(OAc)2 solution. At this point, an exchange of acetate for temperatures resulting in accelerated reaction kinetics. Since the
btc must occur to initiate formation of the MOF structure. Upon step-by-step reactions are slow at room temperature, requiring
exposure to Cu(II) at the second step (Cu(OAc)2/step 2), the several days to grow MOF films with thickness of the order of
metal cation in solution must exchange the acetate anions 100 nm, higher temperatures will enable growth rates that are
coordinated to it (at least partially) with btc bound to the more practical and at temperatures compatible with most
surface. The net effect is again an exchange of organic anions. devices. Second, the HKUST-1 growth rate is dependent on the
One would expect similar activation energies for these reactions, composition of the surface; the order of the growth rates is Al2O3
since the same bond types are being broken in each case, and > SiO2 > OH-SAM > CO2H-SAM. Finally, the morphology of
indeed, the transition state thermodynamics corresponding to k1 the initial substrate has a substantial impact on the quality of the
and k2 are nearly the same. Furthermore, the relatively low DH‡ deposited MOF film, with rougher substrates yielding even
for these reactions suggests that the activated complex involves rougher films.
both Cu(II) and linker species (represented as {Cu(II)/btc}) and The controlled growth of MOFs on surfaces constitutes an
that in {Cu(II)/btc} the bonds are partially broken.45,48 The important step towards developing successful protocols for
solvent likely plays a significant role in stabilizing the transition coating various substrates, with the ultimate goal of integrating
state. MOFs with various functional devices. While many other MOF
We note that it is unclear whether these two processes occur via systems remain to be explored, this investigation illustrates the
a single reaction or a series of individual elementary steps. basic kinetic principles for Cu3(btc)2 film growth on several
Overall, however, the reaction thermodynamics in Table 1 are surfaces relevant to functional device applications and is a stim-
consistent with the fact that Cu3(btc)2 grows quite easily on ulus for similar detailed studies of other MOFs.
surfaces without formation of any significant byproducts. The
model depicted in Fig. 8 is also consistent with previously pub-
lished QCM, SPR and GIXRD results.13,24,42 Recent experi-
Acknowledgements
mental reports suggest that, although the step-by-step growth The authors would like to thank Jeff Chames for his skilful
mechanism is valid for some other MOFs,23,26,34,35,56 it is far from technical assistance. We also would like to thank John J. Perry
being general. In particular, it is likely not applicable to cases in IV for help with the graphical representation of the proposed
which the SBU or an analogue thereof does not exist in solution, growth model and useful discussions. We gratefully acknowledge
as is the case here. For example, IRMOF structures require financial support from Sandia Laboratory Directed Research
formation of a tetrahedral m4-oxo Zn4O SBU that involves and Development Program. Sandia National Laboratories is
significant rearrangement of bonds from those existing in the a multi-program laboratory managed and operated by Sandia
typical precursors, such as Zn(II) nitrate or acetate. In such cases, Corporation, a wholly owned subsidiary of Lockheed Martin
the rate-limiting step may well be a reaction in solution, which Corporation, for the U.S. Department of Energy’s National
would change the observed film growth kinetics and the associ- Nuclear Security Administration under contract DE-AC04-
ated mechanism. In our experience, this is consistent with the 94AL85000.
greater difficulty associated with growing dense films of IRMOF-
1 (MOF-5) on any surface.
References
1 S. M. Cohen, Chem. Sci., 2010, 1, 32–36.
Conclusions 2 D. Zacher, R. Schmid, C. W€ oll and R. A. Fischer, Angew. Chem., Int.
New insights into the step-by-step deposition of Cu3(btc)2 on Ed., 2011, 50, 176–199.
3 O. Shekhah, J. Liu, R. A. Fischer and C. W€ oll, Chem. Soc. Rev., 2011,
silica, alumina and SAM-functionalized Au surfaces were 40, 1081–1106.
obtained using the QCM and QCM-D techniques. In situ 4 D. Zacher, O. Shekhah, C. W€ oll and R. A. Fischer, Chem. Soc. Rev.,
monitoring of MOF growth, coupled with frequency and dissi- 2009, 38, 1418–1429.
5 D. J. Tranchemontagne, J. L. Mendoza-Cortes, M. O’Keeffe and
pation measurements, provides data necessary to establish the O. M. Yaghi, Chem. Soc. Rev., 2009, 38, 1257–1283.
kinetics of film growth and allows calculation of the reaction 6 A. U. Czaja, N. Trukhan and U. Muller, Chem. Soc. Rev., 2009, 38,
rates and transition-state thermodynamics. Using this approach, 1284–1293.
we propose a comprehensive mechanism for the MOF growth on 7 J. Lee, O. K. Farha, J. Roberts, K. A. Scheidt, S. T. Nguyen and
J. T. Hupp, Chem. Soc. Rev., 2009, 38, 1450–1459.
surfaces. The accurate measurement of frequency changes also 8 J. R. Li, R. J. Kuppler and H. C. Zhou, Chem. Soc. Rev., 2009, 38,
allows correlation of the adsorbed mass with the growth condi- 1477–1504.
tions. This case study reveals significant differences in the texture 9 L. E. Kreno, K. Leong, O. K. Farha, M. Allendorf, R. P. Van Duyne
and J. T. Hupp, Chem. Rev., 2012, 112, 1105–1125.
and morphology of the resulting MOF coatings based on the
10 M. D. Allendorf, A. Schwartzberg, V. Stavila and A. A. Talin, Chem.–
substrate used. The temperature-dependent kinetics data show Eur. J., 2011, 17, 11372–11388.
that the activation energy of Cu3(btc)2 thin-film growth is about 11 M. Meilikhov, K. Yusenko, D. Esken, S. Turner, G. Van Tendeloo
an order of magnitude lower than that observed during synthesis and R. A. Fischer, Eur. J. Inorg. Chem., 2010, 3701–3714.
12 A. l. B"etard and R. A. Fischer, Chem. Rev., 2012, 112, 1055–1083.
of bulk Cu3(btc)2 crystals. The magnitudes of the activation 13 O. Shekhah, H. Wang, D. Zacher, R. A. Fischer and C. W€ oll, Angew.
enthalpy and entropy suggest associative reaction mechanisms Chem., Int. Ed., 2009, 48, 5038–5041.
for the transition states corresponding to Cu(II) and btc addition. 14 S. S.-Y. Chui, S. M.-F. Lo, J. P. H. Charmant, A. G. Orpen and
I. D. Williams, Science, 1999, 283, 1148–1150.
This work provides some important practical guidelines for the
15 M. D. Allendorf, R. J. T. Houk, L. Andruszkiewicz, A. A. Talin,
deposition of Cu3(btc)2 on various surfaces. First, HKUST-1 J. Pikarsky, A. Choudhury, K. A. Gall and P. J. Hesketh, J. Am.
film growth process is thermally activated, with higher Chem. Soc., 2008, 130, 14404–14406.

This journal is ª The Royal Society of Chemistry 2012 Chem. Sci., 2012, 3, 1531–1540 | 1539
16 P. Kusgens, M. Rose, I. Senkovska, H. Frode, A. Henschel, S. Siegle 36 R. A. Fischer and C. W€ oll, Angew. Chem., Int. Ed., 2009, 48, 6205–
and S. Kaskel, Microporous Mesoporous Mater., 2009, 120, 325–330. 6208.
17 G. W. Peterson, G. W. Wagner, A. Balboa, J. Mahle, T. Sewell and 37 M. R. Deakin and D. A. Buttry, Anal. Chem., 1989, 61, 1147A–
C. J. Karwacki, J. Phys. Chem. C, 2009, 113, 13906–13917. 1154A.
18 L. S. Grajciar, A. D. Wiersum, P. L. Llewellyn, J.-S. Chang and 38 G. Sauerbrey, Z. Phys., 1959, 155, 206–222.
P. Nachtigall, J. Phys. Chem. C, 2011, 115, 17925–17933. 39 C.-Y. Huang, M. Song, Z.-Y. Gu, H.-F. Wang and X.-P. Yan,
19 H. Wu, J. M. Simmons, Y. Liu, C. M. Brown, X. S. Wang, S. Ma, Environ. Sci. Technol., 2011, 45, 4490–4496.
V. K. Peterson, P. D. Southon, C. J. Kepert, H. C. Zhou, 40 L.-F. Song, C.-H. Jiang, C.-L. Jiao, J. Zhang, L.-X. Sun, F. Xu,
T. Yildirim and W. Zhou, Chem.–Eur. J., 2010, 16, 5205–5214. W.-S. You, Z.-G. Wang and J.-J. Zhao, Cryst. Growth Des., 2010,
20 S. C. Xiang, W. Zhou, J. M. Gallegos, Y. Liu and B. L. Chen, J. Am. 10, 5020–5023.
Chem. Soc., 2009, 131, 12415–12419. 41 O. Zybaylo, O. Shekhah, H. Wang, M. Tafipolsky, R. Schmid,
21 D. Britt, D. Tranchemontagne and O. M. Yaghi, Proc. Natl. Acad. D. Johannsmann and C. W€ oll, Phys. Chem. Chem. Phys., 2010, 12,
Sci. U. S. A., 2008, 105, 11623–11627. 8092–8097.
22 A. L. Robinson, M. Allendorf, V. Stavila and S. M. Thornberg, MRS 42 O. Shekhah, Materials, 2010, 3, 1302–1315.
Online Proc. LIbr., 2011, 1366, uu08-02. 43 K. Szelagowska-Kunstman, P. Cyganik, M. Goryl, D. Zacher,
23 O. Shekhah, H. Wang, T. Strunskus, P. Cyganik, D. Zacher, Z. Puterova, R. A. Fischer and M. Szymonski, J. Am. Chem. Soc.,
R. Fischer and C. W€ oll, Langmuir, 2007, 23, 7440–7442. 2008, 130, 14446–14448.
24 O. Shekhah, H. Wang, S. Kowarik, F. Schreiber, M. Paulus, 44 R. A. Fischer and C. W€oll, Angew. Chem., Int. Ed., 2008, 47, 8164–8168.
M. Tolan, C. Sternemann, F. Evers, D. Zacher, R. A. Fischer and 45 Chemically Reacting Flow: Theory and Practice, ed. R. J. Kee, M. E.
C. W€ oll, J. Am. Chem. Soc., 2007, 129, 15118–15119. Coltrin and G. W. Peterson, Wiley-InterScience, Inc., New Jersey,
25 J. Gascon, S. Aguado and F. Kapteijn, Microporous Mesoporous 2003.
Mater., 2008, 113, 132–138. 46 Encyclopedia of Surface and Colloid Science, ed. P. Somasundaran,
26 D. Zacher, A. Baunemann, S. Hermes and R. A. Fischer, J. Mater. CRC Press, Boca Raton, FL, 2006, vol. 1.
Chem., 2007, 17, 2785–2792. 47 R. W. Tillmann, M. Radmacher and H. E. Gaub, Appl. Phys. Lett.,
27 J. L. Zhuang, D. Ceglarek, S. Pethuraj and A. Terfort, Adv. Funct. 1992, 60, 3111–3113.
Mater., 2011, 21, 1442–1447. 48 Chemical Kinetics and Dynamics, ed. J. I. Steinfeld, J. S. Francisco and
28 A. Schoedel, C. Scherb and T. Bein, Angew. Chem., Int. Ed., 2010, 49, W. L. Hase, Prentice Hall, Inc, New Jersey, 1989.
7225–7228. 49 L. Motiei, M. Feller, G. Evmenenko, P. Dutta and M. E. van der
29 M. Darbandi, H. K. Arslan, O. Shekhah, A. Bashir, A. Birkner and Boom, Chem. Sci., 2012, 3, 66–71.
C. W€ oll, Phys. Status Solidi RRL, 2010, 4, 197–199. 50 N. A. Khan, E. Haque and S. H. Jhung, Phys. Chem. Chem. Phys.,
30 O. Shekhah, H. Wang, M. Paradinas, C. Ocal, B. Schupbach, 2010, 12, 2625–2631.
A. Terfort, D. Zacher, R. A. Fischer and C. W€ oll, Nat. Mater., 51 N. S. John, C. Scherb, M. Shoaee, M. W. Anderson, M. P. Attfield
2009, 8, 481–484. and T. Bein, Chem. Commun., 2009, 6294–6296.
31 C. Munuera, O. Shekhah, H. Wang, C. W€ oll and C. Ocal, Phys. 52 R. E. Morris, ChemPhysChem, 2009, 10, 327–329.
Chem. Chem. Phys., 2008, 10, 7257–7261. 53 M. Shoaee, M. W. Anderson and M. P. Attfield, Angew. Chem., Int.
32 E. Biemmi, A. Darga, N. Stock and T. Bein, Microporous Mesoporous Ed., 2008, 47, 8525–8528.
Mater., 2008, 114, 380–386. 54 M. Shoaee, J. R. Agger, M. W. Anderson and M. P. Attfield,
33 E. Biemmi, C. Scherb and T. Bein, J. Am. Chem. Soc., 2007, 129, 8054. CrystEngComm, 2008, 10, 646–648.
34 O. Shekhah, K. Hirai, H. Wang, H. Uehara, M. Kondo, S. Diring, 55 J. Nan, X. Dong, W. Wang, W. Jin and N. Xu, Langmuir, 2011, 27,
D. Zacher, R. A. Fischer, O. Sakata, S. Kitagawa, S. Furukawa 4309–4312.
and C. W€ oll, Dalton Trans., 2011, 40, 4954–4958. 56 D. Zacher, K. Yusenko, A. Betard, S. Henke, M. Molon, T. Ladnorg,
35 K. Yusenko, M. Meilikhov, D. Zacher, F. Wieland, C. Sternemann, O. Shekhah, B. Schupbach, T. de los Arcos, M. Krasnopolski,
X. Stammer, T. Ladnorg, C. W€ oll and R. A. Fischer, M. Meilikhov, J. Winter, A. Terfort, C. W€ oll and R. A. Fischer,
CrystEngComm, 2010, 12, 2086–2090. Chem.–Eur. J., 2011, 17, 1448–1455.

1540 | Chem. Sci., 2012, 3, 1531–1540 This journal is ª The Royal Society of Chemistry 2012

You might also like