You are on page 1of 12

Article

Cite This: J. Am. Chem. Soc. 2018, 140, 1812−1823 pubs.acs.org/JACS

Crystal Growth of ZIF-8, ZIF-67, and Their Mixed-Metal Derivatives


Daniel Saliba, Manal Ammar, Moustafa Rammal, Mazen Al-Ghoul,* and Mohamad Hmadeh*
Department of Chemistry, American University of Beirut, P.O. Box 11-0236, Riad El-Solh, 1107 2020 Beirut, Lebanon
*
S Supporting Information

ABSTRACT: A facile method to produce zeolitic imidazolate


frameworks (ZIF-8, ZIF-67, and solid−solution ZIFs (mixed
Co and Zn)) is reported. ZIF crystals are produced via a
reaction−diffusion framework (RDF) by diffusing an outer
solution at a relatively high concentration of the 2-methyl
imidazole linker (HmIm) into an agar gel matrix containing
the metal ions (zinc(II) and/or cobalt(II)) at room temper-
Downloaded by UNIV FED DA PARAIBA at 05:28:38:689 on June 06, 2019

ature. Accordingly, a propagating supersaturation wave,


initiated at the interface between the outer solution and the
gel matrix, leads to a precipitation front with a gradient of
crystal sizes ranging between 100 nm and 55 μm along the
from https://pubs.acs.org/doi/10.1021/jacs.7b11589.

reaction tube. While the precipitation fronts of ZIF-8 and ZIF-67 travel the same distance for the same initial conditions, ZIF-8
crystals therein are consistently smaller than the ZIF-67 crystals due to the disparity of their rate of nucleation and growth. The
effects of the temperature, the concentration of the reagents, and the thickness of the gel matrix on the growth of the ZIF crystals
are investigated. We also show that by using RDF we can envisage the formation mechanism of the ZIF crystals, which consists of
the aggregation of ZIF nanospheres to form the ZIF-8 dodecahedrons. Moreover, using RDF, the formation of a solid−solution
ZIF via the incorporation of Co(II) and Zn(II) cations within the same framework is achieved in a controlled manner. Finally, we
demonstrate that doping ZIF-8 by Co(II) enhances the photodegradation of methylene blue dye under visible light irradiation in
the absence of hydrogen peroxide.

■ INTRODUCTION
Zeolitic imidazolate frameworks (ZIFs) are porous crystalline
challenging because of the complexity of adding a second metal
that can lead to the formation of a physical mixture of discrete
materials that consist of tetrahedral clusters of MN4 (e.g., M = structures.24 Very recently, Co-doped ZIF-8 has been used to
Zn(II), Co(II)) linked by simple imidazolate ligands.1−3 ZIFs tweak the size and the physicochemical properties of the
constitute a subfamily of metal−organic frameworks (MOFs) obtained heterometallic structure.25 More importantly, hetero-
and are of great interest for their highly desirable physical and metallic ZIF-8 material exhibits enhanced CO2 and H2 uptakes
chemical properties such as crystallinity, porosity, rich structural and the highest catalytic CO2 conversion as compared to the
diversity, and exceptional chemical and thermal stability.4,5 homometallic ZIF-8 and ZIF-67.11,26 Moreover, ZIF-8 is used
Because of these combined features, ZIFs show significant as a photocatalyst for methylene blue dye degradation.22
potential for gas storage and separation applications.6,7 As in However, it is only active under UV−Vis illumination due to its
the case of zeolites, the properties and performance of porous large optical band gap (Eg = 4.9 eV).27,28 On the other hand,
materials depend on the well-defined size and shape of the while ZIF-67 has a band gap of 1.98 eV, which makes it an
nanocrystals and microcrystals.8 Therefore, it is obvious that excellent visible light absorber, its photocatalytic activity for dye
the new synthetic routes have benefited from the profound degradation is shown to be poor, mainly attributed to its weak
understanding of the physicochemical fundamentals of the chemical stability in aqueous media. As a remedy, Zhang et al.
crystallization processes.9 ZIFs have been synthesized solvo- showed that doping ZIF-67 with Cu(II) enhances chemical
thermally or at room temperature using organic solvents (e.g., stability and thereby improves the visible light photo-
methanol, DMF) or pure water.10,11 The control of size and degradation of methyl orange in the presence of H2O2.22 In a
morphology of crystals has been achieved using reverse very recent report, rhombic dodecahedrons of ZIF-67 were
microemulsion methods,12,13 microwave,14 ultrasound-assisted used as photocatalysts for the reduction of Cr(VI) with an
syntheses,15 and coordination modulation methods.16 In efficiency of around 70% under UV−Vis light irradiation.29
particular, ZIF-8 which possesses a sodalite-type structure is In this paper, we report a new method to synthesize ZIF-8,
the most investigated ZIF material for a variety of applications ZIF-67, and their heterometallic version via a reaction−
including catalysis,17,18 gas sorption,19 and sensing.20,21 One of diffusion framework (RDF) where the 2-methyl imidazolate
the interesting features of this structure is that Zn(II) cations (HmIm)-based solution (outer electrolyte) is added on top of
can be substituted by Co(II) or Cu(II) without altering its
topology.22,23 This can be achieved by postsynthetic treatment Received: October 31, 2017
or via a one-pot synthesis. The latter approach is more Published: January 5, 2018

© 2018 American Chemical Society 1812 DOI: 10.1021/jacs.7b11589


J. Am. Chem. Soc. 2018, 140, 1812−1823
Journal of the American Chemical Society Article

Figure 1. Theoretical basis of the reaction−diffusion framework (RDF). (A) Diffusion profiles of HmIm (outer) and Zn2+/Co2+ (inner) are depicted
at a given time. Evolution of the reaction zone is also shown and exhibits a decrease of its amplitude and broadening of its width. xf denotes the
location of the peak which also corresponds to the location of the precipitation front. (B) Nucleation of pure ZIF-8 leading to nanosperoids takes
place within the reaction zone. Transition to multisized Co-doped polyhedra with a gradient of composition takes place in its wake.

an agar gel matrix containing the coprecipitating metal cations This synthesis of ZIFs via RDF is also advantageous for many
(Zn2+ and/or Co2+), serving as so-called inner electrolytes. In reasons: (i) it is carried out efficiently at room temperature
RDF, because the reaction is taking place in a gel matrix which without the usage of aging processes or thermal treatments, (ii)
hinders the kinetics of precipitation, the mechanism of the it is scalable where gram scales can be obtained by simply using
formation of the solid ZIF material can be portrayed by the larger reactors, (iii) it permits the preparation of solid solution
following consecutive physicochemical processes (Figure 1A): and generation of a spatial doping gradient along the tubular
reactor, and (iv) it easily provides control over the particle size
k
HmIm(aq) + Zn 2 +(aq)/Co2 +(aq) → {ZIF(aq)} → ZIF(s) and morphology.
Moreover, the photocatalytic properties of the obtained
where HmIm diffuses and reacts with Zn2+ and/or Co2+ with a crystals are explored by examining the extent of methylene blue
rate constant k to generate the colloidal intermediate {ZIF(aq)} (MB) degradation by ZIF-8 and the Co(II)-doped versions
in a narrow reaction zone defined by the rate function R = under visible light irradiation without using H2O2.
k[HmIm][M2+] (where M = Zn and/or Co) that exhibits
maximum production rate at location xf which can be used to
define the location of the reaction front. This reaction zone that
■ EXPERIMENTAL SECTION
Materials. 2-Methyl imidazole (HmIm), hydrochloric acid (HCl),
resembles a Gaussian curve is characterized by a time-varying dichloromethane (DCM), zinc and cobalt salts, isopropyl alcohol and
amplitude, width, and speed. Because the chemical reaction is N,N-dimethylformamide (DMF) were purchased from Sigma-Aldrich,
much faster than diffusion, the dynamics of the reaction front is while Bacto Agar was obtained from Difco, and methylene blue (MB)
diffusion controlled and depends mainly on the initial was bought from Acros Organics.
Preparation Method. The inner portion was prepared by adding
concentration of the diffusing molecules and their diffusion
agar powder into an aqueous solution containing the desired amount
coefficients DHmIm, DZn2+, DCo2+, and DZIF.30 The initial of zinc nitrate. The mixture was heated and stirred until the agar
concentrations of the inner and outer electrolytes are chosen dissolved completely; then an equal volume of DMF was added with
such that their product (initial supersaturation) exceeds by far continuous stirring. The obtained solution was poured into a test tube
their solubility product, which in turn leads to a propagating (10 cm × 1 cm or 20 cm × 2.4 cm, depending on the scale of the
supersaturation wave similar in shape and dynamics to the experiment) to fill its two-thirds (5 mL for the smaller tube or 50 mL
reaction zone.31 The diffusing species {ZIF(aq)} which is for the larger one) and then covered with parafilm and left for 2 h at
produced inside the reaction zone will then nucleate if its room temperature to allow complete gelation of agar. After gelation of
the inner portion, the outer solution (HmIm dissolved in a 1:1 mixture
concentration exceeds a certain threshold. These nuclei will of water/DMF) was poured on top of it (2 mL for the smaller tube
subsequently grow into immobile crystals in the wake of the and 10 mL for the larger one). The tube was then covered and left for
reaction zone as depicted in Figure 1B. The gel matrix in RDF 2 days to allow the reaction−diffusion process and formation of the
provides a solid support for the nucleation and growth of the white precipitate front. This process is shown in Scheme 1 and Figure
ZIF crystals, eliminates any disruptive convective currents, and 2A. Interestingly, ZIF crystals can be synthesized using different ratios
slows down the kinetics of the growing crystals. Consequently, of water and DMF. It is noteworthy that in the hydrogel without any
we are able to study the transition from nucleation to DMF the layered ZIF structure (ZIF-L) was obtained (Figure S1).33
The content of the tubes was divided into equidistant bands of
agglomeration leading to the crystal growth of ZIFs in slow about 5 mm with a thickness of 1 mm for each one as shown in Figure
motion (Figure 1B), providing insights into the fundamentals 2B. Each band was extracted from the tube and then washed in DMF
of ZIF crystallization.32 At high supersaturation, especially at until complete redissolution of the gel. After centrifugation, the
the early stages of diffusion near the interface, the rate of crystals were collected for analysis. For the ZIF-67, the zinc nitrate was
nucleation dominates and we obtain small crystals; however, as substituted by cobalt nitrate and a purple precipitate was obtained
the supersaturation weakens far from the interface, growth (Figure 2B). As for the heterometallic ZIF (zinc/cobalt), the same
dominates and we obtain fewer but larger crystals (Figure 1B). preparation method was used where zinc and cobalt nitrate were
mixed in different ratios in the gel. In this case, after reaction−
Hence, the established supersaturation wave leads to a gradient
diffusion, two regions were observed, a relatively narrow leading white
in crystal size and can range from a few nanometers to a few precipitate that gradually transforms into a purple precipitate (Figure
micrometers. This size variation is nonlinear and depends on 2B and movie in SI). For particle size measurements in this case, the
the coupling of the reaction−diffusion process to nucleation solid particles from the leading white band are excluded. For atomic
and growth kinetics. absorbance analysis, consecutive bands were extracted and washed.

1813 DOI: 10.1021/jacs.7b11589


J. Am. Chem. Soc. 2018, 140, 1812−1823
Journal of the American Chemical Society Article

Scheme 1. Reaction Scheme between the Outer Solution of measured and the ratio x = Co(II)/(Co(II) + Zn(II)) was calculated.
HmIm and the Inner Gel solution containing Zn(II) and/or At the macroscopic scale, the same preparation method was used and a
Co(II) To Give the ZIF Crystals thin test tube (0.5 cm × 20 cm), in which the reaction is taking place,
was placed under a Canon EOS 450D camera which was controlled to
take a shot every 60 min for 60 h. The smallness of the tube’s diameter
guarantees a pseudo-one-dimentional setup for front propagation.
Using digital photoanalysis, the distances traveled by each of the ZIF
precipitation fronts, d, are measured as a function of time, t, for
different inner electrolyte concentrations and exhibit a scaling law of
the form d ≈ tn as clearly shown in the log(d) vesus log(t) curves
plotted in Figure 2C for ZIF-8 and ZIF-67. The measured distance can
be considered as xf depicted in Figure 1A, as no reaction occurs
beyond the reaction zone. The exponent values, n, are calculated as the
slopes of the fitted lines and found to be ca. 0.5 for both ZIFs,
confirming that the formation of the ZIFs are indeed diffusion
controlled. This is further endorsed by the fact that both ZIFs travel
the same distance at any given time during propagation for the same
inner and outer concentrations. For such a case, it is mathematically
shown that the position of the front depends only on the initial
concentrations of the inner and outer solutions and their diffusion
coefficients, and since DZn2+ are DCo2+ are almost equal,34 the distance
traveled by the two fronts must be equal.35 Moreover, the lower the
concentration of the inner electrolyte, the faster the propagation of the
ZIF front (Figure 2C). Because the system can be considered as 1D as
mentioned above, the effective diffusion coefficient (Deff) for the three
inner concentrations in Figure 2C can be extracted from the equation
The dry powder was then dissolved in 1 ml of 1 M HCl. After required d2 = Deff × t and is found to be 0.058 (10 mM), 0.051 (25 mM), and
dilutions, molar concentrations of cobalt and zinc cations were 0.039 cm2/s (50 mM).35 It is noteworthy that although both fronts

Figure 2. Schematic representation of the ZIFs synthesis via the reaction−diffusion framework (A). Formation of the ZIF-8, Co-doped ZIF-8, and
ZIF-67 crystals in the tubular reactor where the extracted bands are indicated as bn (B). log−log plots of the displacement of ZIF-8 (blue) and ZIF-
67 (red) fronts for different inner electrolyte concentrations (C).

1814 DOI: 10.1021/jacs.7b11589


J. Am. Chem. Soc. 2018, 140, 1812−1823
Journal of the American Chemical Society Article

Figure 3. PXRD patterns of the simulated ZIF-8/67 (A (i)) and the RDF-synthesized ZIF-8 (A (ii)) and ZIF-67 (A (iii)). Thermogravimetric curves
of ZIF-8 (B (i)) and ZIF-67 (B (ii)). Nitrogen adsorption (solid circle)−desorption (empty circle) isotherms at 77 K of ZIF-8 (C) and ZIF-67 (D).

(ZIF-8 and ZIF-67) travel the same distance under the same
conditions, the particle size distributions for both ZIFs are completely
disparate as we are going to explore in the coming sections.
■ RESULTS AND DISCUSSION
Characterization of the White (ZIF-8) and Purple (ZIF-
ZIFs Characterization. Powder X-ray diffraction (PXRD) patterns 67) Precipitates. The phase purity and structural identity of
were recorded on a Bruker D8 discover XRD diffractometer using Cu the white- and purple-extracted precipitates are determined
Kα radiation (λ = 1.5406 Å) at 40 kV and 40 mA, with 2θ ranging using PXRDs. The recorded patterns exhibit sharp peaks that
between 5° and 40° at a scanning rate of 0.5°·min−1. Thermogravi- match the simulated ones from the sodalite structures of ZIF-8
metric analysis (TGA) of the precipitates was performed under and ZIF-67 (Figure 3A). No extra peaks are observed, which
nitrogen atmosphere with a heating rate of 5 °C·min−1 and a demonstrates the purity of the ZIF samples synthesized using
temperature ranging from 30 to 900 °C using a TG 209 F1 Iris RDF. The observed peak broadening in ZIF-8 crystals suggests
(Netzsch, Germany). The N2 adsorption/desorption isotherms, the formation of nanosized crystals which are also observed by
surface area, and total pore volume were obtained using a high-resolution scanning electron microscopy (HRSEM). The
Micromeritics ASAP 2420 analyzer. Prior to the N2 adsorption/
thermal behavior of the as-synthesized and activated ZIFs is
desorption and TGA measurements, the samples were washed for 3
days with DMF followed by washing with DCM for 2 days. The UV−
examined via thermogravimetric analysis (TGA). Similar to the
Vis diffuse reflectance experiments were recorded on a JASCO V-570 previously reported ZIFs, the TGA curves in Figure 3B
UV−Vis−NIR Spectrophotometer. The morphology of the crystals demonstrate their high thermal stability (up to 500 °C) for
was examined using a scanning electron microscope (Tescan Mira), both ZIFs. To assess the porosity and the architectural stability
operating at 5 kV with an Oxford detector for energy-dispersive X-ray of the ZIF structures, N2 gas adsorption measurements at 77 K
(EDX) characterization. The atomic absorption measurements were are performed. The Brunauer−Emmett−Teller (BET) surface
carried out by a Thermo Labsystems SOLAAR atomic absorption areas for ZIF-8 (Figure 3C) and ZIF-67 (Figure 3D) are
spectrophotometer with ASX-510 autosampler and SOLAAR data calculated to be 1302 and 1023 m2·g−1, respectively. These
acquisition. The photocatalytic activity of the Co-doped ZIF-8 was values agree with the reported values in the literature.25
investigated by degradation of aqueous methylene blue, acting as a Morphological Features and Particle Size Distribution
model pollutant of water, using solar light. The photodegradation of ZIF-8 and ZIF-67. The morphology of the extracted ZIF-8
experiments were carried out in a photoreactor equipped with a 100 W
and ZIF-67 is inspected via HRSEM (Figure 4), which reveals a
xenon lamp, an AM 1.5 optical filter, a magnetic stirrer, and a jacketed
rhombic dodecahedral shape with sharp edges and smooth
beaker, placed 20 cm away from the light source, containing 15 mg of
ZIF powder and 10 mg·L−1 of methylene blue (Vtotal = 50 mL). Prior faces. In order to show the effect of the generated super-
to irradiation, the ZIF/organic dye mixture was stirred in the dark for saturation gradient along the tube, one test tube is prepared at
60 min to reach the adsorption−desorption equilibrium and then room temperature, where the concentrations of the inner
exposed to a solar light source. Samples were taken at time intervals, electrolyte, outer electrolyte, and agar gel are 5 mM, 2 M, and
and the residual methylene blue was detected by a spectrophotometer 1% (w/w), respectively. After 48 h of propagation, six
at λmax = 665 nm. equidistant bands (Figure 2B), having a thickness of 1 mm
1815 DOI: 10.1021/jacs.7b11589
J. Am. Chem. Soc. 2018, 140, 1812−1823
Journal of the American Chemical Society Article

Figure 4. Particle size distributions of ZIF-8 (A) and ZIF-67 (B) extracted from six consecutive equidistant (5 mm) bands along the tube. Each band
has a thickness of 1 mm.

Figure 5. Variation of the average ZIF-8 (A) and ZIF-67 (B) particle size with respect to the band distance from the gel/outer electrolyte interface.
Different band numbers are indicated in Figure 2B.

and a separation of 5 mm, are extracted and examined by electrolyte), a supersaturation wave is generated by the
HRSEM. The size of the dodecahedrons gradually increases interdiffusing species along the tube starting at the gel/outer
from 500 nm near the interface (band 1 (b1)) until it reaches 4 interface downward. Due to diffusion, this supersaturation wave
μm at the white leading front (band 6 (b6)) for ZIF-8 (Figures is characterized by a slowly decreasing amplitude and widening
4A and 5A) and from 1 to 10 μm for ZIF-67 (Figures 4B and breadth.36−38 Because the rates of nucleation and growth of the
5B). Upon addition of the 2-methylimidazole solution (outer ZIF particles strongly and nonlinearly depend on the
1816 DOI: 10.1021/jacs.7b11589
J. Am. Chem. Soc. 2018, 140, 1812−1823
Journal of the American Chemical Society Article

Figure 6. Effect of temperature (A), concentration of the inner electrolyte (B), and concentration of the agar gel (C) on the average particle size
distribution of the ZIF-8 crystals. Each band is approximately equal to 1 mm, and the consecutive bands are equidistant (5 mm), as indicated in
Figure 2B. Average ZIFs size is reported as a mean of a sample size of 20 particles, and the error bars represent the standard deviation.

magnitude of the supersaturation, we obtain a gradient of sizes size. In a second set of reactions, three test tubes are prepared
along the tubular reactor, although the extracted bands are in which the zinc nitrate concentration is varied (10, 50, and
equidistant. Near the interface, nucleation dominates and 100 mM), while the concentrations of the outer electrolyte and
smaller particles are formed; conversely, as we go farther from gel are kept fixed at 1 M and 1% (w/w), respectively. Near to
the interface, growth starts to dominate and bigger particles are the interface (band 1), the particle size increases as the inner
obtained. Interestingly, in the first bands near the interface we concentration increases: (1) 0.12 μm, 10 mM; (2) 0.95 μm, 50
are able to observe the structural preorganization of the mM; (3) 3.1 μm, 100 mM. The same trend is obtained as we
nanocrystals before formation of the rhombic dodecahedral- go farther from the interface (Band 2: (1) 1.7 μm, 10 mM; (2)
shaped crystals of ZIF-8 and ZIF-67 (Figure 4). The faceted 2.3 μm, 50 mM; (3) 4.1 μm, 100 mM. Band 3: (1) 2.6 μm, 10
particles appear far from the gel/outer interface. Such features mM; (2) 4.1 μm, 50 mM. Band 4:(1) 3.2 μm, 10 mM; (2) 5.0
of crystal growth cannot be observed in conventional methods μm, 50 mM), where smaller particles are obtained for lower
of ZIF synthesis, which makes RDF a very useful and unique inner concentrations, as shown in Figure 6B. As the metal ion
method to capture the different growth stages of the crystals. (inner) concentration increases, at a fixed imidazole (outer)
Consequently, RDF provides a synthetic tool that easily allows concentration, the generated initial supersaturation gradient
manipulation of size distributions through alteration of will be smaller and the obtained particles are bigger when
supersaturation by simply varying different experimental compared at the same distance from the gel/outer interface.
parameters, such as the temperature of the reaction medium, Effect of the Gel. The effect of the agar gel concentration
the inner electrolyte types and concentrations, and the on the ZIF-8 crystals size distribution is also investigated. Three
thickness of the gel. The effect of these parameters is different experiments are carried out at room temperature and
investigated for ZIF-8 crystal synthesis. at a gel concentration of 0.5%, 1%, and 1.5%, where the
Effect of Temperature. Four different experiments are concentrations of the inner and outer electrolytes are kept fixed
carried out for 48 h at different temperatures, 7, 15, 20, and 25 at 100 mM and 1 M, respectively. From the three bands
°C, where the concentrations of the inner electrolyte (zinc extracted from each tube we observe that near the interface
nitrate), outer electrolyte, and gel are kept fixed at 50 mM, 1 M, (band 1) the particle size decreases as the gel thickness
and 1% (w/w), respectively. Near to the interface (band 1), the increases: (1) 1.6 μm, 0.5%; (2) 1.0 μm, 1%; (3) 0.51 μm,
particle size increases with the temperature: (1) 0.3 μm, 7 °C; 1.5%. The same trend is obtained as we go farther from the
(2) 0.6 μm, 15 °C; (3) 0.7 μm, 20 °C; (4) 0.8 μm, 25 °C. The interface (Band 2: (1) 7.5 μm, 0.5%; (2) 5.9 μm, 1%; (3) 4.4
same trend is obtained as we go farther from the interface μm, 1.5%. Band 3: (1) 10.1 μm, 0.5%; (2) 7.7 μm, 1%; (3) 6.3
(Band 2: (1) 0.8 μm, 7 °C; (2) 1.9 μm, 15 °C; (3) 2.9 μm, 20 μm, 1.5%), where smaller particles are obtained for thicker gel,
°C; (4) 3.6 μm, 25 °C. Band 3: (1) 1.1 μm, 7 °C; (2) 2.6 μm, as shown in Figure 6C. The inversely proportional relation
15 °C; (3) 3.8 μm, 20 °C; (4) 4.2 μm, 25 °C), where bigger between the gel concentration and the particle size is due to the
particles are obtained as the temperature increases (Figures 6A generation of smaller pores as the gel thickness increases.37 By
and S2). This trend in particle size over all of the extracted decreasing the pore size, the growth of the crystals is hindered
bands is a direct consequence of the increase of the rate of and smaller particles are obtained.
growth with temperature as we move away along the Nucleation and Growth Kinetics of ZIFs Dodecahe-
precipitation front. drons. The nucleation and growth kinetics of the ZIF crystals
Effect of Inner Electrolyte. The effect of the inner at their inception is investigated. For this study, ZIF-8 and ZIF-
electrolyte is also investigated. First, the type of metal salt is 67 are prepared in 2 identical reactors with 20 mM inner and 1
varied, and four different metal sources for zinc and cobalt are M outer concentrations. The crystals are then extracted from
used (acetate, chloride, sulfate, and nitrate). SEM (Figure S3) the gel/outer interface, where the highest supersaturation is
measurements confirm the phase purity of ZIF-8 and ZIF-67. encountered in each reactor at different times (30, 60, and 120
The size distributions obtained from the different reactions are s) by discarding the outer electrolyte. PXRDs of the solids at
shown in Figure S4. While the usage of different zinc salts does the first minute correspond to the ZIF-8 and ZIF-67 structures
not reveal any significant effect on the particle size distribution and indicate that the formation of the ZIF crystals is very fast
of ZIF-8, the effect on the size of ZIF-67 crystals is noticeable (Figure S5). The SEM investigation of the gel/outer electrolyte
with those derived from the acetate having relatively the biggest interface region reveals that the formation of ZIF-8 and ZIF-67
1817 DOI: 10.1021/jacs.7b11589
J. Am. Chem. Soc. 2018, 140, 1812−1823
Journal of the American Chemical Society Article

Figure 7. Nucleation and growth kinetics of ZIFs dodecahedrons. SEM images of the ZIF-8 crystals after 30 (A), 60 (B), and 120 s (C) and of ZIF-
67 crystals after 30 (D), 60 (E), and 120 s (F) of reaction at the interface region.

Figure 8. Cobalt doping of ZIF-8 using RDF for different initial (Co(II)/(Co(II) + Zn(II)) ratios exhibiting a decreasing gradient of cobalt content
as we move away from the outer electrolyte/gel interface (A) along the reaction tube. Visible diffuse reflectance spectra of ZIF-8 (black line), ZIF-67
(blue line), and CoxZn1−x(HmIm)2 (red line) (B).

crystals is preceded by an instantaneous homogeneous diffusion process (Figure 5). This growth mechanism is
nucleation leading to nanospheroids with an average diameter consistent with the one reported in the literature.9,27,39
of less than 100 nm (Figure 7A and 7D). However, the Doping with Cobalt and Growth Mechanism of ZIF-8.
appearance of faceted crystals of ZIF-8 (Figure 7B) was slower It has been demonstrated that Co(II) cations substitute Zn(II)
than that of ZIF-67 (Figure 7E and 7F) as evidenced by the without altering the SOD topology of the ZIF structure.22,29 In
SEM images. The aggregation of these nanospheres, probably order to apply this doping process in RDF, different
via oriented attachment, leads to the formation of faceted percentages of Co(II) are added to the inner Zn(II) electrolyte.
crystals within 120 s for ZIF-8 (Figure 7C) and 60 s for ZIF-67
Interestingly, upon addition of the outer electrolyte (HmIm)
(Figure 7E).27 The size of resulting ZIF-8 crystals (ca. 100 nm)
two regions are observed: a leading white precipitate that
at the interface is smaller than that of ZIF-67 (ca. 300 nm).
This can be explained by the fact that ZIF-67 is not water stable gradually transforms into a purple one (Figure 8A). Therefore,
as compared to ZIF-8, and consequently, its rate of nucleation the incorporation of Co(II) into the ZIF-8 structure to form
is expectedly smaller because fewer nuclei survive thermal the purple CoxZn1−x(HmIm)2 can be monitored via a visible
fluctuations. Hence, nucleation for ZIF-8 dominates the crystal color change of the precipitate. The UV−Vis diffuse reflectance
formation, whereas growth governs in the case of ZIF-67, spectrum, recorded for the purple precipitate, exhibits
leading to noticeably larger crystals throughout the reaction− characteristic bands of tetrahedral-coordinated Co(II) at 585
1818 DOI: 10.1021/jacs.7b11589
J. Am. Chem. Soc. 2018, 140, 1812−1823
Journal of the American Chemical Society Article

Figure 9. SEM image of a Co-doped ZIF-8 dodecahedron and its corresponding elemental mapping of Co and Zn.

and 540 nm (Figure 8B), confirming the incorporation of CoxZn1−x(HmIm)2 ranges between 0.33 (in band 1) and
Co(II) in the ZIF-8 structure. 0.01(in band 4). Moreover, the maximum precentage of doped
The PXRD pattern for the purple precipitate is identical to cobalt is found in the first bands and represents 50% to 75% of
the ZIF-8 pattern, which indicates that the synthesized Co- the initial concentration ratio and decreases gradually to reach
doped ZIF-8 exhibits a SOD topology identical to the parent less than 30% in the leading bands. From this set of doping
framework. Since Co(II) and Zn(II) have slightly different ionic experiments, we conclude that the incorporation of Zn is
radii (0.74 Å for Zn(II) and 0.72 Å for Co(II)), it makes it thermodynamically more favored than that of Co in the ZIF
difficult to uncover any shift toward lower angles for the Co- framework.
doped samples (Figure S6). For the surface composition The measured particle sizes for the heterometallic
analysis, an energy-dispersive X-ray (EDX) analysis is CoxZn1−x(HmIm)2 crystals are consistently smaller than both
conducted on the purple precipitate and reveals the coexistence the ZIF-8 and the ZIF-67 ones in all bands (Figures 11 and S7).
of the elements Zn and Co in the same crystal, confirming the
formation of a solid solution of ZIF-8 and ZIF-67 and not a
physical mixture of both (Figure 9). To gain more insight into
the doping reaction mechanism, 10 different initial cobalt
percentages are used in the inner electrolytes (5%, 10%, 20%,
30%, 40%, 50%, 60%, 70%, 80%, and 90%). The different tubes
are divided into 4 equidistant bands (both the band separation
and the band thickness are equal to 5 mm) to determine the
homogeneity of the substitution throughout the precipitation
region. The crystals in each of the 4 bands are extracted,
washed with DMF, and examined by EDX and atomic
absorption. Figure 10 shows the atomic percentage of cobalt
in the ZIF-8 crystal structure for the four bands in each of the
10 experiments. Remarkably, the percentage of cobalt
incorporated in the structure gradually decreases as the distance
from the gel/outer interface increases in all tubes. For example,
Figure 11. Variation of the particle size of the ZIF-8 dodecahedrons as
for an initial mixture of 1:1 Co(II):Zn(II), the value of x in
a function of the initial Co(II) percentage present in the inner
electrolyte and the distance from the gel/outer electrolyte interface.
Each band has a thickness of 5 mm, and each of the two consecutive
bands are separated by 5 mm.

This is contrary to what Zareba et al.25 observed for their


heterometallic Co−Zn−ZIFs, where they found that the
particle size increases exponentially with the Co percentage.
Their particle size range was between 20 and 500 nm; however,
in our RDF system, the obtained particle sizes fall in a different
range (from 100 nm to 10 μm). This trend in crystal size of the
CoxZn1−x(HmIm)2 crystals is seemingly due to Co(II) ions that
act as “impurities” with respect to ZIF-8 crystal nucleation that
Figure 10. Variation of the Co(II) incorporation percentage in the starts first as evidenced by the leading white front. Therefore, it
ZIF-8 crystal structure as a function of the Co(II) percentage present looks like a plausible idea that the presence of cobalt enhances
in the inner electrolyte and the distance from the gel/outer interface. the kinetics of nucleation, leading to doped particles of
Each band has a thickness of 5 mm, and each of the two consecutive relatively smaller size. This is further endorsed by a higher
bands are separated by 5 mm. initial concentration of cobalt ions reaching a 95% ratio (Figure
1819 DOI: 10.1021/jacs.7b11589
J. Am. Chem. Soc. 2018, 140, 1812−1823
Journal of the American Chemical Society Article

S8). Interestingly, however, at 100% cobalt (pure ZIF-67), the temperature, the local depletion rate in the zinc cations
size of the obtained crystal is distinctively bigger. becomes lower; thus, less cobalt is consumed for growth of the
The effect of the temperature on the doping reaction is particles.
further investigated, and three different temperatures (7, 23, From the SEM images of the doped samples, we were able to
and 30 °C) are chosen with an initial Co(II) concentration of capture the structural evolution of the solid−solution ZIFs. A
10%. The efficiency of Co doping is increased by a factor of 2 reorganization of nanospheroidal particles in the white leading
for an increase of the reaction temperature from 7 to 30 °C front to form the faceted particles in the purple precipitation
(Figure 12). This enhancement in the percentage of Co zone can be observed in Figure 13. This comportment was
observed for all of the doped samples as shown, for example, in
the SEM images (Figure S9) for the white front region for three
different Co-doping percentages (1%, 5%, and 10%). Here the
white front is further partitioned into three consecutive bands
(Figure S9A−C) to capture the agglomeration of the
nanospheres (Figure 9A) into polyhedral-shaped structures
(Figure 9B) before assuming well-defined single crystals
(Figure 9C). We then selected the 1% doping case for PXRD
analysis (Figure S10) because it relatively provides the widest
white front (ca. 0.5 cm) with enough powder for analysis. The
lowest band in the white front away from the white−pink
interface is composed of nanospheroids (50−100 nm) of
crystalline pure ZIF-8 (from EDX data) and exhibits broad-
ening of the peaks in the PXRD data (Figure S10A). This
indicates that in addition to its higher thermodynamic stability,
ZIF-8 is also kinetically more favored than the Co-doped ZIF-8.
Figure 12. Effect of temperature on the cobalt-doping percentage of Furthermore, as the agglomeration of the white spheres starts
the ZIF-8 crystals. near the white−purple interface in the second band of the
white front, the doping begins to take place and the PXRD
incorporation is observed for all bands in the tube. As peaks sharpen (Figure S10B) until they reach the final cobalt
mentioned above, the ZIF-8 formation is kinetically more composition, size, and well-defined polyhedral morphology in
favored, and hence, the growth of the ZIF-8 dodecahedrons band 3 of the white front located at the white−purple interface
brings a local depletion in the concentration of Zn(II) at the and above (Figure S10C). Moreover, the SEM images in the
expense of Co(II), which will then be incorporated in the transition region in Figure 13 show the existence of hollow
structure. Since smaller particles are obtained at lower assembly. We believe that some of the formed polyhedral might

Figure 13. SEM images of the leading white front (A), transition zone from white to purple (B−E), and upper purple precipitate (F) from a Co-
doped ZIF-8 system showing the formation mechanism of the ZIF dodecahedrons.

1820 DOI: 10.1021/jacs.7b11589


J. Am. Chem. Soc. 2018, 140, 1812−1823
Journal of the American Chemical Society Article

remain hollow and others are fully filled to form single crystals. former and 3 min for the latter (Figure S16). Remarkably, the
Self-assembly in MOFs and ZIFs crystals to form hollow Co-doped ZIF-8 exhibits higher dye adsorption (ca. 65%) than
macrostructures is common and was recently reviewed.40 This that of ZIF-8 (ca. 40%) at saturation under our experimental
feature deserves further investigation in the future. This relative conditions. This notable difference in saturation time and
ease in monitoring crystal formation constitutes one of the adsorption capacity between the two solids is one of the
main advantages of using RDF and confirms the mechanism important facets of doping.
obtained using MD simulations.41 Kinetically, the UV−vis spectra of the MB solutions exhibit a
Stability of Co-Doped ZIF-8. Because ZIF-67 is not water diminution as a function of chromophoric absorbance
stable, the chemical stability of Co-doped ZIF-8 is investigated. maximum of MB centered at 665 nm42 as shown in Figure
We extracted doped ZIF-8 crystals with various cobalt S17. Interestingly, 70% of the MB is photodegraded in 90 min
concentrations and washed them with water for 24 h. when the Co-doped ZIF-8 (10%) is used, whereas only 45% is
Remarkably, the doped products for an initial Co(II) degraded when ZIF-8 is used (Figure 14). The kinetic data in
concentration below 50% are chemically stable as evidenced Figure 14 are fitted to a stretched exponential, and the rate
from their PXRDs (Figure S11). Nevertheless, for high initial constants in the case of the pure and Co-doped ZIF-8 are
Co(II) concentrations (70% and above), the obtained solid− extracted (Table S2) and found to differ, in the first cycle, by 1
solution ZIFs are not chemically stable. order of magnitude (0.0044 ± 0.0011 and 0.021 ± 0.001 min−1
Photocatalytic Properties of ZIF Crystals. The gen- for ZIF-8 and the Co-doped ZIF-8, respectively). On the other
erated ZIF crystals are tested for the photocatalytic degradation hand, the high chemical stability of the samples after the
of methylene blue dye under visible light irradiation. The band photodegradation reactions is confirmed as evidenced by the
gaps (Eg) of ZIF-8, ZIF-67, and the heterometallic ZIF are PXRD and SEM shown in Figure S18. Moreover, the
determined from their diffuse reflectance spectra (converted reusability of both photocatalysts is tested, and remarkably,
using Kubelka−Munk function) (Figure 8B) and found to be both ZIFs retained their relative activity in the second cycle
4.9, 2.0, and 2.0 eV for ZIF-8, ZIF-67, and Co-doped ZIF-8 with rate constants of 0.0021 ± 0.001 min−1 for ZIF-8 and
(10%), respectively. This indicates that our doping approach is 0.016 ± 0.001 min−1 for Co-doped ZIF-8 (Figure 14). On the
useful to control the optical properties and to bring the ZIF-8 other hand, it is noteworthy that since ZIF-67 is not chemically
material to the realm of visible light absorption and stable in water, it cannot be used in such applications. It is also
photocatalysis. In order to test for any enhancement in important to mention that the reported dye photodegradation
photocatalytic performance of ZIF-8, ZIF-67, and Co-doped studies using ZIF materials employed H2O2 to produce •OH
ZIF-8, we evaluate their activity during the degradation of MB radicals that speed up the oxidation process22 or strong UV
under visible light illumination without using H2O2.22 The rate light irradiation,43 whereas our experiments are performed in
and extent of MB degradation are determined from the time- aqueous solution in the absence of H2O2 and under visible light
dependent decrease of the MB concentration, which is illumination.
calculated from the decrease in the visible absorbance band Under visible light illumination, electrons are excited from
(665 nm) of MB. It is clear from Figures S12 and S13 that the the HOMO to the LUMO of the ZIF material, generating holes
performance of all of the doped ZIF is similar. Therefore, we and electrons. The photogenerated electrons react with the
choose the 10% Co-doped ZIF-8 (surface area of 1235 m2·g−1 dissolved or adsorbed oxygen on the surface of ZIF crystals to
(Figure S14)) to compare with the pure ZIF-8. The control produce the superoxide radical anion •O2−, while the holes
solution, consisting of pure MB with no catalyst, shows a small react with water to form reactive hydroxyl radicals •OH, which
percentage of degradation (around 10%) after 90 min of probably cause the degradation of MB as deduced from the
illumination (Figures 14 and S15). Moreover, in the absence of peaks around 210 nm in the case of ZIF-8 and around 272−280
light, the MB solution in the presence of ZIF-8 and the 10% nm for the Co-doped ZIF-8 (Figure S17). Moreover, only in
Co-doped-ZIF-8 displayed a saturation period of 45 min for the the case of ZIF-8, the aforementioned reaction competes with
the decolorization of MB by reduction to yield leuco methylene
blue (LMB), which is usually observed at 256 nm43,44 (Figure
S17A and S17B). The mechanism involving •OH is further
confirmed by performing the same photodegradation reaction
in the presence of isopropyl alcohol as hole scavenger. In this
experiment, we performed photodegradation tests with the 10%
Co-doped ZIF-8. In this case the degradation was almost
completely inhibited as shown in Figure S19. A similar
mechanism had been proposed recently for the degradation
of organic dyes in the presence of metal−organic frame-
works.45,46 Under similar experimental conditions, Jelle et al.
recently reported the photodegradation of MB over FeOOH
materials and showed that only 20% of MB is degraded after 90
min, and to reach 70% of degradation more than 300 min is
needed.42 Their turnover frequency, TOF (TOF = 2.2 × 10−4
min−1), is found to be relatively lower than ours for the 10%
Co-doped ZIF-8 (TOF = 3.0 × 10−4 min−1). In addition to the
enhancement of visible light absorption by the doping strategy,
Figure 14. Photocatalytic performance of the ZIF-8 and 10% Co- the efficiency of our catalyst can also be attributed to the
dopedZIF-8 in the degradation of methylene blue (MB) under visible remarkably enhanced adsorption of MB on the surface of the
light and for two consecutive cycles. doped material, as mentioned above, which is a favored
1821 DOI: 10.1021/jacs.7b11589
J. Am. Chem. Soc. 2018, 140, 1812−1823
Journal of the American Chemical Society Article

precursor for heterogeneous catalysis. The obtained results


demonstrate the importance of Co doping on enhancing the
■ REFERENCES
(1) Eddaoudi, M.; Kim, J.; Rosi, N.; Vodak, D.; Wachter, J.; O’keeffe,
photocatalytic properties of ZIF-8. M.; Yaghi, O. M. Science 2002, 295, 469−472.

■ CONCLUSIONS
We synthesized and fully characterized ZIF-8 and ZIF-67
(2) Rosi, N. L.; Eckert, J.; Eddaoudi, M.; Vodak, D. T.; Kim, J.;
O’keeffe, M.; Yaghi, O. M. Science 2003, 300, 1127−1129.
(3) Furukawa, H.; Ko, N.; Go, Y. B.; Aratani, N.; Choi, S. B.; Choi,
E.; Yazaydin, A. Ö .; Snurr, R. Q.; O’Keeffe, M.; Kim, J.; Yaghi, O. M.
crystals using a reaction−diffusion framework (RDF) where the Science 2010, 329, 424−428.
precipitation reaction resulted from the interdiffusion of initially (4) Chen, R.; Yao, J.; Gu, Q.; Smeets, S.; Baerlocher, C.; Gu, H.; Zhu,
separated metal ions and organic linkers in agar gel as a matrix. D.; Morris, W.; Yaghi, O. M.; Wang, H. Chem. Commun. 2013, 49,
The supersaturation gradient in RDF provided us a means to 9500−9502.
effectively control the size of the ZIF crystals. Crystal size (5) Tang, J.; Salunkhe, R. R.; Liu, J.; Torad, N. L.; Imura, M.;
distributions were also studied as a function of the Furukawa, S.; Yamauchi, Y. J. Am. Chem. Soc. 2015, 137, 1572−1580.
concentration of the inner electrolytes, the temperature of (6) Hu, Y.; Liu, Z.; Xu, J.; Huang, Y.; Song, Y. J. Am. Chem. Soc. 2013,
the reaction medium, and the thickness of the gel. 135, 9287−9290.
(7) Rodenas, T.; Luz, I.; Prieto, G.; Seoane, B.; Miro, H.; Corma, A.;
Consequently, we demonstrated that this method is effective
Kapteijn, F.; Llabres i Xamena, F. X. L.; Gascon, J. Nat. Mater. 2015,
in tuning with ease the crystal size distribution at ambient 14, 48−55.
temperature. We also studied the doping of ZIF-8 by cobalt (8) Tosheva, L.; Valtchev, V. P. Chem. Mater. 2005, 17, 2494−2513.
ions by means of RDF. This also revealed a gradient of doping (9) Venna, S. R.; Jasinski, J. B.; Carreon, M. A. J. Am. Chem. Soc.
concentration along the tubular reaction reactor. Furthermore, 2010, 132, 18030−18033.
we took advantage of the slow rates of nucleation and growth (10) Jian, M.; Liu, B.; Liu, R.; Qu, J.; Wang, H.; Zhang, X. RSC Adv.
inside the gel pores to visualize the growth mechanism of the 2015, 5, 48433−48441.
ZIF-8 crystals and their Co-doped analogues that consisted of (11) Kaur, G.; Rai, R. K.; Tyagi, D.; Yao, X.; Li, P.-Z.; Yang, X.-C.;
an aggregation of ZIF-8 nanospheres to form well-resolved Zhao, Y.; Xu, Q.; Singh, S. K. J. Mater. Chem. A 2016, 4, 14932−
polyhedra. Finally, the photocatalytic properties of the Co- 14938.
(12) Sun, W.; Zhai, X.; Zhao, L. Chem. Eng. J. 2016, 289, 59−64.
doped ZIF-8 crystals were evaluated by studying the
(13) Shang, W.; Kang, X.; Ning, H.; Zhang, J.; Zhang, X.; Wu, Z.;
degradation of MB under visible light irradiation. We Mo, G.; Xing, X.; Han, B. Langmuir 2013, 29, 13168−13174.
demonstrated the importance of Co doping in tuning the (14) Bao, Q.; Lou, Y.; Xing, T.; Chen, J. Inorg. Chem. Commun. 2013,
optical and photocatalytic activity of ZIF-8 for dye degradation 37, 170−173.
under visible light irradiation. (15) Safarifard, V.; Morsali, A. Coord. Chem. Rev. 2015, 292, 1−14.


(16) Guo, H.; Zhu, Y.; Wang, S.; Su, S.; Zhou, L.; Zhang, H. Chem.
ASSOCIATED CONTENT Mater. 2012, 24, 444−450.
(17) Chizallet, C.; Lazare, S.; Bazer-Bachi, D.; Bonnier, F.; Lecocq,
* Supporting Information
S V.; Soyer, E.; Quoineaud, A.-A.; Bats, N. J. Am. Chem. Soc. 2010, 132,
The Supporting Information is available free of charge on the 12365−12377.
ACS Publications website at DOI: 10.1021/jacs.7b11589. (18) Kuo, C.-H.; Tang, Y.; Chou, L.-Y.; Sneed, B. T.; Brodsky, C. N.;
Zhao, Z.; Tsung, C.-K. J. Am. Chem. Soc. 2012, 134, 14345−14348.
Characterization of the ZIF-L crystals; SEMs, BET, (19) Banerjee, R.; Furukawa, H.; Britt, D.; Knobler, C.; O’Keeffe, M.;
particle size distribution of the ZIF crystals for different Yaghi, O. M. J. Am. Chem. Soc. 2009, 131, 3875−3877.
counter anions, stability of the Co-doped ZIF-8 in (20) Lu, G.; Hupp, J. T. J. Am. Chem. Soc. 2010, 132, 7832−7833.
aqueous solution, and full spectra of the photo- (21) Liu, S.; Xiang, Z.; Hu, Z.; Zheng, X.; Cao, D. J. Mater. Chem.
degradation of MB using different ZIFs (PDF) 2011, 21, 6649−6653.
Movie for the front propagation and formation of Co- (22) Yang, H.; He, X.-W.; Wang, F.; Kang, Y.; Zhang, J. J. Mater.
Chem. 2012, 22, 21849−21851.
doped ZIF-8 in the reaction-diffusion framework. (23) Tang, J.; Salunkhe, R. R.; Zhang, H.; Malgras, V.; Ahamad, T.;
(MOV) Alshehri, S. M.; Kobayashi, N.; Tominaka, S.; Ide, Y.; Kim, J. H. Sci.


Rep. 2016, 6, 30295.
AUTHOR INFORMATION (24) Zhou, K.; Mousavi, B.; Luo, Z.; Phatanasri, S.; Chaemchuen, S.;
Verpoort, F. J. Mater. Chem. A 2017, 5, 952−957.
Corresponding Authors (25) Zaręba, J. K.; Nyk, M.; Samoć, M. Cryst. Growth Des. 2016, 16,
*mazen.ghoul@aub.edu.lb 6419−425.
*mohamad.hmadeh@aub.edu.lb (26) Kuruppathparambil, R. R.; Babu, R.; Jeong, H. M.; Hwang, G.-
Y.; Jeong, G. S.; Kim, M.-I.; Kim, D.-W.; Park, D.-W. Green Chem.
ORCID 2016, 18, 6349−6356.
Mohamad Hmadeh: 0000-0003-3027-3192 (27) Cravillon, J.; Schröder, C. A.; Nayuk, R.; Gummel, J.; Huber, K.;
Notes Wiebcke, M. Angew. Chem. 2011, 123, 8217−8221.
(28) Wang, C.-C.; Li, J.-R.; Lv, X.-L.; Zhang, Y.-Q.; Guo, G. Energy
The authors declare no competing financial interest.


Environ. Sci. 2014, 7, 2831−2867.
(29) Park, H.; Amaranatha Reddy, D.; Kim, Y.; Ma, R.; Choi, J.; Kim,
ACKNOWLEDGMENTS T. K.; Lee, K.-S. Solid State Sci. 2016, 62, 82.
(30) Abi Mansour, A.; Al Ghoul, M. Phys. Rev. E 2011, 84, 026107.
The authors gratefully acknowledge the funding provided by
(31) Al-Ghoul, M.; Ammar, M.; Al-Kaysi, R. O. J. Phys. Chem. A
the American University of Beirut Research Board and by the 2012, 116, 4427−4437.
Lebanese National Council for Scientific Research (LCNSR). (32) Al-Ghoul, M.; Issa, R.; Hmadeh, M. CrystEngComm 2017, 19,
M.H. acknwoledges funds from The World Academy of Science 608−612.
(TWAS) grant #103277. We also acknowledge the help of Mr. (33) Low, Z.-X.; Razmjou, A.; Wang, K.; Gray, S.; Duke, M.; Wang,
Antranik Jonderian in the elemental mapping experiment. H. J. Membr. Sci. 2014, 460, 9−17.

1822 DOI: 10.1021/jacs.7b11589


J. Am. Chem. Soc. 2018, 140, 1812−1823
Journal of the American Chemical Society Article

(34) Lide, D. R.; Kehiaian, H. V.; CRC Handbook of Thermophysical


and Thermochemical Data; CRC Press: Boca Raton, FL, 1994.
(35) Gálfi, L.; Rácz, Z. Phys. Rev. A: At., Mol., Opt. Phys. 1988, 38,
3151.
(36) Saliba, D.; Ezzeddine, A.; Sougrat, R.; Khashab, N. M.; Hmadeh,
M.; Al-Ghoul, M. ChemSusChem 2016, 9, 800−805.
(37) Saliba, D.; Al-Ghoul, M. CrystEngComm 2016, 18, 8445−8453.
(38) Saliba, D.; Ezzeddine, A.; Emwas, A.-H.; Khashab, N. M.; Al-
Ghoul, M. Cryst. Growth Des. 2016, 16, 4327−4335.
(39) Seoane, B.; Castellanos, S.; Dikhtiarenko, A.; Kapteijn, F.;
Gascon, J. Coord. Chem. Rev. 2016, 307, 147−187.
(40) Furukawa, S.; Reboul, J.; Diring, S.; Sumida, K.; Kitagawa, S.
Chem. Soc. Rev. 2014, 43, 5700−5734.
(41) Pan, Y.; Heryadi, D.; Zhou, F.; Zhao, L.; Lestari, G.; Su, H.; Lai,
Z. CrystEngComm 2011, 13, 6937−6940.
(42) Jelle, A. A.; Hmadeh, M.; O’Brien, P. G.; Perovic, D. D.; Ozin,
G. A. ChemNanoMat 2016, 2, 1047−1054.
(43) Pegu, R.; Majumdar, K. J.; Talukdar, D. J.; Pratihar, S. RSC Adv.
2014, 4, 33446−33456.
(44) Impert, O.; Katafias, A.; Kita, P.; Mills, A.; Pietkiewicz-Graczyk,
A.; Wrzeszcz, G. Dalton Trans. 2003, 348−353.
(45) Jing, H.-P.; Wang, C.-C.; Zhang, Y.-W.; Wang, P.; Li, R. RSC
Adv. 2014, 4, 54454−54462.
(46) Wang, C.-c.; Jing, H.-p.; Wang, P. J. Mol. Struct. 2014, 1074, 92−
99.

1823 DOI: 10.1021/jacs.7b11589


J. Am. Chem. Soc. 2018, 140, 1812−1823

You might also like