scispace - formally typeset
Open AccessJournal ArticleDOI

Metal−Organic Frameworks as Sensors: A ZIF-8 Based Fabry−Pérot Device as a Selective Sensor for Chemical Vapors and Gases

Guang Lu, +1 more
- 20 May 2010 - 
- Vol. 132, Iss: 23, pp 7832-7833
Reads0
Chats0
TLDR
A ZIF-8 thin film-based Fabry-Pérot device has been fabricated as a selective sensor for chemical vapors and gases as well as a series of ZIF -8 thin films grown on silicon substrates.
Abstract
A ZIF-8 thin film-based Fabry−Perot device has been fabricated as a selective sensor for chemical vapors and gases. The preparation of the ZIF-8 thin film and a series of ZIF-8 thin films of various thicknesses grown on silicon substrates are presented.

read more

Content maybe subject to copyright    Report

Metal-Organic Frameworks as Sensors: A ZIF-8 Based Fabry-Pe´rot Device
as a Selective Sensor for Chemical Vapors and Gases
Guang Lu and Joseph T. Hupp*
Department of Chemistry, Northwestern UniVersity, EVanston, Illinois 60208
Received February 17, 2010; E-mail: j-hupp@northwestern.edu
Permanently microporous metal-organic framework (MOF)
materials
1
have attracted considerable attention on account of their
large internal surface areas, uniform channels, (sub)nanometer-sized
cavities, thermal stability, and chemical tailorability.
2
Among the
potential applications for which proof-of-concept reports have
proliferated are gas storage,
3
chemical catalysis,
4
and small-
molecule separations.
5
At first glance MOFs would also appear to
be attractive for chemical sensing. Nevertheless, reports of MOF-
based sensing are comparatively few.
6
The challenge is signal
transduction: the cavities of MOFs are generally too small to tailor
with reporter molecules, i.e. moieties that can readily signal analyte-
binding events via changes in color, redox potential, or other
properties. In the handful of cases where sensing has been described,
advantage has generally been taken of framework luminescence,
7
with signal transduction consisting of luminescence quenching.
6,8
Here we describe an alternative approach that circumvents the
need for molecular-level reporters and instead relies upon a readout
of changes in a macroscopic property of the sensing material, the
refractive index. Specifically, we have configured the material as a
transparent thin film on an appropriate support material (glass or
silicon) and optically monitored the energies of Fabry-Pe´rot
interference peaks as a function of analyte exposure.
9
These peaks
are observable when the thickness of a supported material is
comparable to the wavelength of light (λ). Importantly, their
energies also depend on the film’s refractive index.
As the sensing material we chose a zeolitic imidazolic frame-
work, ZIF-8. This chemically robust and thermally stable material
is characterized by the sodalite (SOD) zeolite-type structure with
large cavities (11.6 Å) and small pore apertures (3.4 Å).
10
Diverse
protocols
10
are available to produce ZIF-8 crystals with sizes from
micrometers to nanometers.
10d
While dense ZIF-8 membranes have
recently been synthesized on porous supports using solvothermal
conditions for the purpose of gas separation,
11
it is not easy to
precisely control thicknesses of films on the submicrometer scale
using these conditions. Several strategies have been developed to
prepare MOF thin films,
12
but they either produce poor morphol-
ogies
13
or are time-consuming
14
to grow with thicknesses over 100
nm. In our experiments, ZIF-8 was obtained in thin-film form
(Figure 1A) simply by immersing glass or silicon slides in a freshly
prepared methanolic solution of 2-methylimidazole and Zn(NO
3
)
2
at room temperature (see Supporting Information (SI)). Quartz
crystal microgravimetry experiments (gold surfaces; see SI) show
initially rapid film growth that subsequently slows and is largely
complete after 30 min. Thicker films can be obtained by repeating
the process with fresh solutions (Figure 1B and 1C). Cross-sectional
scanning electron microscopy (SEM) images demonstrate that the
films thicken by ca. 100 nm per growth cycle (Figure 1D). (The
initial cycle, however, yields a ca. 50 nm thick film.) X-ray
diffraction measurements (see SI) establish that the films comprise
ZIF-8.
Figure 2 shows a series of ZIF-8 films of various thicknesses, l,
grown on silicon substrates. While ZIF-8 itself is colorless, the
silicon-supported thin films are not. The variable coloration is a
manifestation of film-thickness-dependent optical interference in
the visible region. At normal incidence the wavelengths of the
(reflected) interference peaks are given by
where m is an integer and n is the thin film’s refractive index.
The key to chemical sensing is the tunability of n. For a
microporous MOF, n is a volume-weighted average of the indices
for the cavities (vacuum, n
vac
) 1) and the framework (n
fram
> 1).
Inserting into the cavity any polarizable molecule will displace
vacuum and increase the overall refractive index. From eq 1, this
will result in red shifts of the interference peaks.
Fabry-Pe´rot fringes for transmission of visible light through a
ca. 1000 nm thick film (10-cycle growth) of ZIF-8 on glass were
monitored to demonstrate the sensing properties.
15
Exposure to
propane shifts the visible-region fringes by up to 49 nm (Figure
3A). The extents of the shifts are sensitive to the propane partial
pressure (Figure 3B). The analyte-induced shift occurs within 1
Figure 1. Cross-sectional SEM images of ZIF-8 films grown on silicon
substrates with cycles of (A) 1, (B) 10, and (C) 40; (D) thickness of ZIF-8
film versus numbers of growth cycles.
Figure 2. Photograph of a series of ZIF-8 films of various thicknesses
grown on silicon substrates.
mλ ) 2nl (1)
10.1021/ja101415b XXXX American Chemical Society J. AM. CHEM. SOC. XXXX, xxx, 000
9
A

min and is reversed by replacing analyte with pure nitrogen gas
(see SI). From the magnitudes of the shifts in the interference
fringes, we calculated that the volume fraction of analyte in the
framework is 0.13 for pure propane at 1 atm.
While responsive to a range of vapors and gases, the ZIF-8 sensor
does display some chemical selectivity. For example, linear
n-hexane is readily sensed, but the sterically more demanding
cyclohexane is not, consistent with the small portal size for ZIF-8
cavities. Consistent with the known hydrophobicity of ZIF-8,
10a
the sensor is unresponsive to water vapor. In contrast, ethanol is
readily detected (Figure 4A). Exposure to the vapor above ethanol/
water mixtures of various ethanol contents gives rise to ethanol-
concentration-dependent responses, with the sensor response satu-
rating at ca. 40% ethanol (Figure 4B). We have taken advantage
of this observation to calculate the volume fraction of analyte in
the framework: 0.25 for pure ethanol, which agrees reasonably well
with the void volume fraction (0.20) indicated by single-crystal
X-ray structural measurements.
16
The detection limit depends, in part, on the resolution of the
spectrophotometer. For 1 nm resolution, the estimated detection
limit for ethanol in water is ca. 0.3 vol %, corresponding to an
ethanol vapor concentration of ca. 100 ppm.
In summary, we have constructed MOF-based Fabry-Pe´rot
devices that function as selective sensors for chemical vapors and
gases. To obtain suitable devices we rely upon a simple, but
controllable and effective, method for fabricating ZIF-8 films.
Virtues of this method include gentle reaction conditions (room
temperature), rapid growth rate (100 nm/30 min), good control
over thickness ( 100 nm/cycle), no special requirements for
surface modification of substrates, and ease of removal of solvent
(methanol).
Acknowledgment. We thank DTRA, AFOSR, and the North-
western Nanoscale Science and Engineering Center for support of
our work.
Supporting Information Available: Materials and methods, QCM,
XRD, EDXS, FT-IR, TGA, N
2
gas sorption, gas- and vapor-testing
data, and effective refractive index and volume fraction calculation.
This material is available free of charge via the Internet at http://
pubs.acs.org.
References
(1) (a) Tranchemontagne, D. J.; Mendoza-Corte´s, J. L.; O’Keeffe, M.; Yaghi,
O. M. Chem. Soc. ReV. 2009, 38, 1257–1283. (b) Yaghi, O. M.; O’Keeffe,
M.; Ockwig, N. W.; Chae, H. K.; Eddaoudi, M.; Kim, J. Nature 2003,
423, 705–714. (c) Fere´y, G. Chem. Soc. ReV. 2008, 37, 191–214. (e)
Kitagawa, S.; Kitaura, R.; Noro, S. Angew. Chem., Int. Ed. 2004, 43, 2334–
2375.
(2) For a recent review, see: Wang, Z.; Cohen, S. M. Chem. Soc. ReV. 2009,
38, 1315–1329.
(3) For recent reviews, see: (a) Murray, L. J.; Dincaˇ, M.; Long, J. R. Chem.
Soc. ReV. 2009, 38, 1294–1314. (b) Collins, D. J.; Zhou, H.-C. J. Mater.
Chem. 2007, 17, 3154–3160.
(4) For recent reviews, see: (a) Lee, J. Y.; Farha, O. K.; Roberts, J.; Scheidt,
K. A.; Nguyen, S. T.; Hupp, J. T. Chem. Soc. ReV. 2009, 38, 1450–1459.
(b) Ma, L.; Abney, C.; Lin, W. Chem. Soc. ReV. 2009, 38, 1248–1256.
(5) For a recent review, see: Li, J.-R.; Kuppler, R. J.; Zhou, H.-C. Chem. Soc.
ReV. 2009, 38, 1477–1504.
(6) (a) Chen, B. L.; Yang, Y.; Zapata, F.; Lin, G. N.; Qian, G. D.; Lobkovsky,
E. B. AdV. Mater. 2007, 19, 1693–1696. (b) Chen, B. L.; Wang, L. B.;
Zapata, F.; Qian, G. D.; Lobkovsky, E. B. J. Am. Chem. Soc. 2008, 130,
6718–6719. (c) Chen, B. L.; Wang, L. B.; Xiao, Y. Q.; Fronczek, F. R.;
Xue, M.; Cui, Y. J.; Qian, G. D. Angew. Chem., Int. Ed. 2009, 48, 500–
503. (d) Harbuzaru, B. V.; Corma, A.; Rey, F.; Jorda´, J. L.; Ananias, D.;
Carlos, L. D.; Rocha, J. Angew. Chem., Int. Ed. 2009, 48, 6476–6479. (e)
Harbuzaru, B. V.; Corma, A.; Rey, F.; Atienzar, P.; Jordá, J. L.; García,
H.; Ananias, D.; Carlos, L. D.; Rocha, J. Angew. Chem., Int. Ed. 2008, 47,
1080–1083. (f) Rieter, W. J.; Taylor, K. M. L.; Lin, W. J. Am. Chem. Soc.
2007, 129, 9852–9853. (g) Xie, Z.; Ma, L.; deKrafft, K. E.; Jin, A.; Lin,
W. J. Am. Chem. Soc. 2010, 132, 922–923. (h) Wong, K. L.; Law, G. L.;
Yang, Y. Y.; Wong, W. T. AdV. Mater. 2006, 18, 1051–1054. (i) Zhao,
B.; Chen, X. Y.; Cheng, P.; Liao, D. Z.; Yan, S. P.; Jiang, Z. H. J. Am.
Chem. Soc. 2004, 126, 15394–15395.
(7) For a recent review, see: Allendorf, M. D.; Bauer, C. A.; Bhakta, R. K.;
Houk, R. J. T. Chem. Soc. ReV. 2009, 38, 1330–1352.
(8) For an exception, see: Allendorf, M. D.; Houk, R. J.; Andruszkiewicz, L.;
Talin, A. A.; Pikarsky, J.; Choudhury, A.; Gall, K. A.; Hesketh, P. J. J. Am.
Chem. Soc. 2008, 130, 14404–14405.
(9) (a) Lin, V. S.-Y.; Motesharei, K.; Dancil, K.-P. S.; Sailor, M. J.; Ghadiri,
M. R. Science 1997, 278, 840–843. (b) Pan, S. L.; Rothberg, L. J. Nano
Lett. 2003, 3, 811–814. (c) Li, Y. Y.; Cunin, F.; Link, J. R.; Gao, T.; Betts,
R. E.; Reiver, S. H.; Chin, V.; Bhatia, S. N.; Sailor, M. J. Science 2003,
299, 2045–2047.
(10) (a) Park, K. S.; Ni, Z.; Coˇte´, A. P.; Choi, J. Y.; Huang, R.; Uribe-Romo,
F. J.; Chae, H. K.; O’Keeffe, M.; Yaghi, O. M. Proc. Natl. Acad. Sci.
U.S.A 2006, 103, 10186–10191. (b) Huang, X.-C.; Lin, Y.-Y.; Zhang, J.-
P.; Chen, X.-M. Angew. Chem., Int. Ed. 2006, 45, 1557–1559. (c) Li, K.;
Olson, D. H.; Seidel, J.; Emge, T. J.; Gong, H.; Zeng, H.; Li, J. J. Am.
Chem. Soc. 2009, 131, 10368–10369. (d) Cravillon, J.; Mu¨nzer, S.;
Lohmeier, S.-J.; Feldhoff, A.; Huber, K.; Wiebcke, M. Chem. Mater. 2009,
21, 1410–1412.
(11) (a) Venna, S. R.; Carreon, M. A. J. Am. Chem. Soc. 2010, 132, 76–78. (b)
Bux, H.; Liang, F.; Li, Y.; Cravillon, J.; Wiebcke, M.; Caro, J. J. Am.
Chem. Soc. 2009, 131, 16000–16001.
(12) (a) Zacher, D.; Shekhah, O.; Wo¨ll, C.; Fischer, R. A. Chem. Soc. ReV.
2009, 38, 1418–1429. (b) Demessence, A.; Horcajada, P.; Serre, C.;
Boissie`re, C.; Grosso, D.; Sanchez, C.; Fe´rey, G. Chem. Commun. 2009,
7149–7151.
(13) (a) Hermes, S.; Schröder, F.; Chelmowski, R.; Wo¨ll, C.; Fischer, R. A.
J. Am. Chem. Soc. 2005, 127, 13744–13745. (b) Biemmi, E.; Scherb, C.;
Bein, T. J. Am. Chem. Soc. 2007, 129, 8054–8055.
(14) (a) Shekhah, O.; Wang, H.; Kowarik, S.; Schreiber, F.; Paulus, M.; Tolan,
M.; Sternemann, C.; Evers, F.; Zacher, D.; Fischer, R. A.; Wo¨ll, C. J. Am.
Chem. Soc. 2007, 129, 15118–15119. (b) Shekhah, O.; Wang, H.; Zacher,
D.; Fischer, R. A.; Wo¨ll, C. Angew. Chem., Int. Ed. 2009, 48, 5038–5041.
(c) Shekhah, O.; Wang, H.; Paradinas, M.; Ocal, C.; Schu¨pbach, B.; Terfort,
A.; Zacher, D.; Fischer, R. A.; Wo¨ll, C. Nat. Mater. 2009, 8, 481–484. (d)
Kanaizuka, K.; Haruki, R.; Sakata, O.; Yoshimoto, M.; Akita, Y.; Kitagawa,
H. J. Am. Chem. Soc. 2008, 130, 15778–15779.
(15) From the interference fringes, the refractive index of evacuated ZIF-8 is
1.39.
(16) The accessible void volume fraction in ZIF-8 framework is estimated in
the CALC SOLV routine in PLATON using a default value of 1.20 Å for
the probe radius.
JA101415B
Figure 3. (A) UV-vis transmission spectra of 10-cycle ZIF-8 film grown
on glass substrate after exposure to propane of various concentrations (blue
curve for 0% and red curve for 100%) and (B) corresponding interference
peak (originally at 612 nm) shift versus propane concentration. The propane
concentration is expressed as a percentage of the total gas flow where
nitrogen is used as diluent.
Figure 4. (A) UV-vis transmission spectra of 10-cycle ZIF-8 film grown
on glass substrate on exposure to vapors of ethanol and water. (B)
Interference peak (originally at 612 nm) shift versus ethanol concentration
in ethanol/water solutions. The concentration is expressed as a volume
percentage.
B J. AM. CHEM. SOC.
9
VOL. xxx, NO. xx, XXXX
COMMUNICATIONS
Citations
More filters
Journal ArticleDOI

Metal–Organic Framework Materials as Chemical Sensors

TL;DR: The potential to computationally predict, with good accuracy, affinities of guests for host frameworks points to the prospect of routinely predesigning frameworks to deliver desired properties.
Journal ArticleDOI

Luminescent Metal–Organic Frameworks

TL;DR: This critical review discusses the origins of MOF luminosity, which include the linker, the coordinated metal ions, antenna effects, excimer and exciplex formation, and guest molecules.
Journal ArticleDOI

Metal-organic frameworks: functional luminescent and photonic materials for sensing applications.

TL;DR: This comprehensive review summarizes the topical developments in the field of luminescent MOF and MOF-based photonic crystals/thin film sensory materials.
References
More filters
Journal ArticleDOI

Functional porous coordination polymers.

TL;DR: The aim is to present the state of the art chemistry and physics of and in the micropores of porous coordination polymers, and the next generation of porous functions based on dynamic crystal transformations caused by guest molecules or physical stimuli.
Journal ArticleDOI

Reticular synthesis and the design of new materials

TL;DR: This work has shown that highly porous frameworks held together by strong metal–oxygen–carbon bonds and with exceptionally large surface area and capacity for gas storage have been prepared and their pore metrics systematically varied and functionalized.
Journal ArticleDOI

Selective gas adsorption and separation in metal–organic frameworks

TL;DR: This critical review starts with a brief introduction to gas separation and purification based on selective adsorption, followed by a review of gas selective adsorbents in rigid and flexible MOFs, and primary relationships between adsorptive properties and framework features are analyzed.
Journal ArticleDOI

Metal–organic framework materials as catalysts

TL;DR: A critical review of the emerging field of MOF-based catalysis is presented and examples of catalysis by homogeneous catalysts incorporated as framework struts or cavity modifiers are presented.
Journal ArticleDOI

Exceptional chemical and thermal stability of zeolitic imidazolate frameworks

TL;DR: Study of the gas adsorption and thermal and chemical stability of two prototypical members, ZIF-8 and -11, demonstrated their permanent porosity, high thermal stability, and remarkable chemical resistance to boiling alkaline water and organic solvents.
Related Papers (5)
Frequently Asked Questions (7)
Q1. What contributions have the authors mentioned in the paper "Metal-organic frameworks as sensors: a zif-8 based fabry-pérot device as a selective sensor for chemical vapors and gases" ?

Among the potential applications for which proof-of-concept reports have proliferated are gas storage, chemical catalysis, and smallmolecule separations. The challenge is signal transduction: the cavities of MOFs are generally too small to tailor with reporter molecules, i. e. moieties that can readily signal analytebinding events via changes in color, redox potential, or other properties. Here the authors describe an alternative approach that circumvents the need for molecular-level reporters and instead relies upon a readout of changes in a macroscopic property of the sensing material, the refractive index. 

Virtues of this method include gentle reaction conditions (room temperature), rapid growth rate (∼100 nm/30 min), good control over thickness (∼ 100 nm/cycle), no special requirements for surface modification of substrates, and ease of removal of solvent (methanol). 

From the magnitudes of the shifts in the interference fringes, the authors calculated that the volume fraction of analyte in the framework is ∼0.13 for pure propane at 1 atm. 

For 1 nm resolution, the estimated detection limit for ethanol in water is ca. 0.3 vol %, corresponding to an ethanol vapor concentration of ca. 100 ppm. 

Exposure to the vapor above ethanol/ water mixtures of various ethanol contents gives rise to ethanolconcentration-dependent responses, with the sensor response saturating at ca. 40% ethanol (Figure 4B). 

The authors have taken advantage of this observation to calculate the volume fraction of analyte in the framework: 0.25 for pure ethanol, which agrees reasonably well with the void volume fraction (0.20) indicated by single-crystal X-ray structural measurements. 

Fabry-Pérot fringes for transmission of visible light through a ca. 1000 nm thick film (10-cycle growth) of ZIF-8 on glass were monitored to demonstrate the sensing properties.