Вы находитесь на странице: 1из 26

Please cite this article in press as: Koo et al.

, Metal-Organic Frameworks for Chemiresistive Sensors, Chem (2019), https://doi.org/10.1016/


j.chempr.2019.04.013

Review
Metal-Organic Frameworks
for Chemiresistive Sensors
Won-Tae Koo,1,2 Ji-Soo Jang,1,2 and Il-Doo Kim1,2,*

Highly sensitive and selective chemical sensors are needed for use in a wide The Bigger Picture
range of applications such as environmental toxic gas monitoring, disease diag- Metal-organic frameworks (MOFs)
nosis, and food quality control. Although some chemiresistive sensors have have attracted much attention in
been commercialized, grand challenges still remain: ppb-level sensitivity, accu- diverse research communities
rate cross-selectivity, and long-term stability. Metal-organic frameworks because of their ultrahigh surface
(MOFs) with record-breaking surface areas and ultrahigh porosity are ideal areas, high porosity, and tunable
sensing materials because chemical sensors rely highly on surface reactions. In structures. In particular, MOFs are
addition, MOFs can be used as a membrane to utilize their unique gas adsorp- considered one of the most ideal
tion and separation characteristics. Furthermore, the use of MOFs as precursors sensing materials since chemical
to enable facile production of various nanostructures is further combined with sensing properties are mainly
other functional materials. Based on these fascinating features of MOFs, there influenced by surface reactions.
have been great efforts to elucidate reaction mechanisms and address limita- Recently, the use of MOFs in
tions in MOF-based chemiresistors. In this review, we present a comprehensive chemiresistive sensors that
overview and recent progress in chemiresistive sensors developed by using transduce electrical signals from
pure MOFs, MOF membranes, and MOF derivatives. surface reactions has rapidly
emerged. The development of
INTRODUCTION conductive MOFs has fueled the
Superior chemical sensors are needed to monitor environmental conditions, to use of pure MOFs as a new class of
realize an early diagnosis of disease, and to control food quality.1–4 With the techno- chemiresistors. MOFs with unique
logical advances in the fields of nanomaterials, the Internet of Things, and artificial gas adsorption and separation
intelligence, the studies on chemical sensors have undergone exceptional growth properties also enable their use in
in recent years.5 For the realization of such innovative technologies in real life, gas sensors as selective filtration
ever more sensitive and selective gas-sensing devices with outstanding responses layers. In addition, as sacrificial
particularly toward part-per-million (ppm) or part-per-billion (ppb) levels of target templates, MOFs can be
gases are required. Thus far, carbon-based nanomaterials and semiconducting converted to various types of gas-
metal oxides (SMOs) have been extensively investigated for the detection of target sensitive nanomaterials such as
gases. However, there are still grand challenges in the performance of chemical sen- carbon composites and metal
sors in terms of sensitivity, selectivity, responding speed, and stability. For example, oxides via controlled pyrolysis or
metal oxides are typically operated at an elevated temperature (200 C‒400 C),6 calcination. In this review, we
which often lead to poor selectivity and low stability due to the high reactivity of summarize the latest studies on
metal oxides with interfering gases at high temperatures. In general, the operation MOF-based chemiresistive
of chemical sensors relies on the transduction of signals such as optical, mechanical, sensors and suggest future
or electrical changes induced by the surface reaction of the analytes. Thus, in order research directions.
to achieve superior sensing performances, an accelerated surface reaction is essen-
tial, transducing the sensing signals more efficiently.

Metal-organic frameworks (MOFs), consisting of metal nodes and organic linkers,


have exceptionally high surface area, ultrahigh porosity, and diverse structures.7
Inspired by these fascinating features, MOFs have been widely explored in various
fields, such as gas storage and separation,8,9 energy applications,10 catalysis,11,12
and chemical sensors.13 In particular, the use of MOFs in chemical gas sensors has
been intensively investigated in an effort to utilize their extraordinary properties

Chem 5, 1–26, August 8, 2019 ª 2019 Elsevier Inc. 1


Please cite this article in press as: Koo et al., Metal-Organic Frameworks for Chemiresistive Sensors, Chem (2019), https://doi.org/10.1016/
j.chempr.2019.04.013

Figure 1. Schematic Illustration of Pure MOFs, MOF Membranes, and MOF Derivatives for
Chemiresistive Sensors

that can promote the surface reactions and adsorptions of gas molecules. Lumines-
cence,14 localized surface plasmon resonance,15 interferometry,16 and electrome-
chanical-based17 MOF sensors have been introduced to transduce gas-sensing sig-
nals in the last several years. These MOF-based chemical sensors showed interesting
and promising sensing properties. However, their sensing systems are rather
complicated, and further enhancement in the transduction values of sensing signals
is required. For instance, MOFs with a low band gap or high electrical conductivity
can effectively transfer the electrical signals. In addition, unique gas storage and
separation properties of MOFs enable selective chemical sensors. Moreover, the
extensive post-synthetic methods of MOFs have opened up a variety of possibil-
ities,18 thus diverse MOF structures with controlled compositions can be easily ob-
tained via MOF-templating routes. Inspired by these innovative characteristics of
MOFs, the researches on MOF-based chemiresistive sensors, which are operated
by a change in electrical resistance following the surface reactions or adsorptions
of target gases, have rapidly grown in recent years. However, there have been no
focused review papers for introducing a comprehensive understanding on the cur-
rent status, relevant challenges, and limitations of pure MOFs, MOF membranes,
and MOFs derivatives for chemiresistive sensors, although important review papers
related to MOF-based chemiresistors have been reported recently.19–23

In this review, we present recent progress on various MOFs, particularly optimized


for applications in chemiresistive sensors. First, we summarize the basic principles
and key parameters of chemiresistive sensors and describe the challenges of con-
1Department of Materials Science and
ventional gas-sensing materials and the opportunity of MOF-based chemiresistors.
Engineering, Korea Advanced Institute of Science
Then, we highlight the representative articles focused on potential uses of MOFs for and Technology, Daejeon 34141, Republic of
applications in chemiresistors: (1) pure MOFs, including three-dimensional (3D) Korea
2Advanced Nanosensor Research Center, KI
MOFs, and two-dimensional (2D) conductive MOFs as a new class of sensing mate-
Nanocentury, Korea Advanced Institute of
rials; (2) MOF membranes, as selective gas filtration layers, on conventional sensing Science and Technology, Daejeon 34141,
materials; and (3) MOF derivatives, such as carbon composites and metal oxides, for Republic of Korea
highly sensitive and selective chemiresistors (Figure 1). The various synthetic routes *Correspondence: idkim@kaist.ac.kr
and sensing characteristics of 3D MOFs, 2D conductive MOFs, MOF-metal https://doi.org/10.1016/j.chempr.2019.04.013

2 Chem 5, 1–26, August 8, 2019


Please cite this article in press as: Koo et al., Metal-Organic Frameworks for Chemiresistive Sensors, Chem (2019), https://doi.org/10.1016/
j.chempr.2019.04.013

composites, MOF-metal oxides composites, MOF-derived carbon composites, and


MOF-derived metal oxides are discussed. Moreover, we emphasize the important
correlation between MOFs-driven material properties and sensing characteristics
and discuss the prevailing challenges as well as the future perspective. We hope
that this review inspires scientists in various fields and guides the development of
next-generation chemical sensors.

BASIC PRINCIPLES AND CHALLENGES OF CHEMIRESISTIVE SENSORS


Basic Principles of Chemiresistive Sensors
The sensing mechanism of chemiresistive sensors is mainly attributed to the transfer
of electrons or holes caused by the surface reactions or adsorptions of gas molecules
on the sensing materials.6 Therefore, when sensing layers react with the adsorbed
target gases, the resistance (or conductance) change of the chemiresistive sensors
occurs. Apart from the simple sensing mechanism, chemiresistive sensors offer
several benefits, including low-cost fabrication, facile integration with various elec-
tronic devices, and ease of miniaturization. The origin of the change in the electrical
resistance in chemiresistive sensors depends on the type of sensing material.
Because of the diversity and high compatibility of MOFs and MOF derivatives with
other materials, multiple principles can be incorporated for MOF-based chemiresis-
tors. For instance, the adsorbed analytes on the surface of pure MOFs donate
electrons to pure MOFs or deprive them of their electrons, causing the change in
the resistance of the sensors.24 The metal nodes and active functional organic groups
in MOFs can act as effective adsorption sites that are beneficial to chemical sensors.
The redox reaction of transition metals in MOFs can also affect the conductivity of
pure MOFs.25 In addition, the volume change of MOFs upon gas adsorption modu-
lates the number of electrons hopping between MOFs, causing resistance changes.26

In general, the sensing mechanisms of composite layers are predominantly gov-


erned by majority phases. Thus, sensing operations of MOF derivatives and MOF
composites are also determined by major components, which can be selected
from metal oxides, carbon composites, nitrides, sulfides, and so on. Gas sensors us-
ing metal oxides, which are the well-known sensing materials, are highly dependent
on the surface reactions between target gas molecules and chemisorbed oxygen
species (O2, O, and O2).6 At an elevated temperature, oxygen molecules in air
are chemisorbed on the surfaces of metal oxides by trapping electrons in metal ox-
ides, generating either electron depletion regions for the case of n-type SMOs or
hole accumulation layers for p-type SMOs. Reducing or oxidizing gas molecules
react with the chemisorbed oxygen species adsorbed on SMOs, modulating the
thickness of electron depletion regions or hole accumulation layers, resulting in
the resistance change of the sensors. On the other hand, the resistance changes
of carbon-based materials, such as carbon nanotubes (CNTs), graphene, graphene
oxides, carbonized MOFs, and so on, are mainly caused by the adsorption of target
analytes on their surfaces.27,28 Depending on the position of adsorption sites, the
adsorbed gas molecules (1) directly interact with carbon-based materials by
donating electrons or depriving electrons (intra-carbon) or (2) change the electron
hopping currents between the sensing materials (inter-carbon) by swelling. In addi-
tion, the sensing behaviors of 2D transitional metal dichalcogenides (MoS2, WS2,
and SnS2) are mainly attributed to the direct interaction between target analytes
and sensing materials,29 similar to the intra-carbon-sensing mechanism.

Based on these mechanisms, we briefly summarize the important parameters of sen-


sors in terms of response (or sensitivity), selectivity, speed (response and recovery

Chem 5, 1–26, August 8, 2019 3


Please cite this article in press as: Koo et al., Metal-Organic Frameworks for Chemiresistive Sensors, Chem (2019), https://doi.org/10.1016/
j.chempr.2019.04.013

times), and stability. We will further discuss the key points for developing highly sen-
sitive and selective MOF-based sensors.

Response and Sensitivity


The response (R) of a sensor is defined as a ratio of resistance (or conductance)
change to baseline resistance (or conductance).27

R (for resistance change) = (Rgas ‒ R0)/R0 = DR/R0 (Equation 1)

R (for conductance change) = (Ggas ‒ G0)/G0 = DG/G0, (Equation 2)

where R0 or Go is the baseline resistance or conductance, and Rgas or Ggas is the resis-
tance or conductance of a sensor when exposed to target analytes. In addition, the
sensitivity (S), which is defined as the slope of the resistance change (dy) of sensors
versus the concentration change (dx) of gas molecules,30 can be also used as a
parameter for evaluating gas-sensing capability:

S = dy/dx. (Equation 3)

Since the response of chemiresistors highly relies on the surface reaction or adsorp-
tion of a target gas on the sensors, large surface areas and high reactivity to the
target gas are needed to achieve a higher response. The high response can also
lead to a low detection limit, because the limit of detection (LOD) is determined
as the lowest concentration at which the response is 3-fold higher than the standard
deviation (snoise) of the baseline resistance (R0)31:

LOD = 3 3 rmsnoise/S, (Equation 4)

where rmsnoise is the root-mean-square noise of the sensors and S is the sensitivity
(dy/dx).

Response and Recovery Times


The response and recovery times are used to quantify the sensing speed of sensors.
In general, the response time is defined as the time taken for the resistance to in-
crease from the baseline resistance (R0) to 90% of the maximum resistance change
(R0 + 0.9DRmax) at a given level of a target gas.30 On the other hand, the recovery
time is defined as the time required for the resistance to decrease from the maximum
resistance change (R0 + DRmax) to 10% of the maximum resistance change (R0 +
0.1DRmax).30 The response and recovery times of conductance-based chemiresistors
are also defined in the same way, except that the baseline conductance (G0) and the
maximum conductance change (DGmax) are used. To develop fast responding sen-
sors, it is necessary to lower the activation energy of surface reactions and accelerate
adsorptions of analytes by using catalysts.

Selectivity and Stability


Selectivity is one of the most important sensing parameters because there is a variety
of interfering gas molecules, which is detrimental to accurate target gas detection.
Selectivity of sensors is usually investigated by comparing the cross-sensitivity to-
ward various analytes at given concentrations.32 If there is a unique surface reaction
with specific analytes, the sensors can have superior selectivity. The selectivity is
hugely affected by various factors such as operating temperatures, composition of
the sensors, humidity, and reactivity between target molecules and sensing

4 Chem 5, 1–26, August 8, 2019


Please cite this article in press as: Koo et al., Metal-Organic Frameworks for Chemiresistive Sensors, Chem (2019), https://doi.org/10.1016/
j.chempr.2019.04.013

materials at a given operating temperature. The decoration of selectors or sensi-


tizers (catalysts) on the sensing materials can promote specific surface reactions
with the target gas, leading to improved selectivity. In addition, sensor stability is
an important parameter. The long-term operation of sensors with high reliability is
critical for their practical application. Typically, when the sensing materials are
exposed to water vapors in air, water vapors are adsorbed on the active sites of
sensing layers, leading to the deterioration of the sensing properties.6

Challenges of Conventional Materials and Opportunities of MOFs


Several challenges of present chemiresistive sensors include low sensitivity (or
response), poor selectivity, and instability. Although metal oxides have relatively
higher response than other materials, they have drawbacks such as low selectivity
and baseline drift caused by high temperature operation.33 Room temperature
operating carbon-based materials, even though they possess relatively high surface
area, suffer from low response, poor selectivity, and low reproducibility.34 In addi-
tion, 2D transition metal dichalcogenides (TMDs) have critical issues; they are easily
oxidized in air.29 To date, a number of researchers have attempted to address
several critical issues of above-mentioned conventional sensing materials, but
further significant improvements are still needed.

MOFs, as new emerging materials, have various fascinating features for use in chem-
iresistive gas sensors (Figure 1). First, MOFs have giant surface areas with high gas
accessibility compared with conventional porous materials.7 If the entire surface of
a MOF is fully utilized for transduction of the sensing signals (resistance changes),
one can expect a dramatic enhancement of the gas response. Second, the pristine
MOFs with high structural tunability have been considerably studied as efficient
membrane layers for gas separation or gas storage.8 These MOFs with selective
gas penetration and adsorption properties offer a potential solution to address
the selectivity issue in chemiresistive sensors. Last, MOF derivatives, which are easily
obtained from the post-synthetic processes of pure MOFs, still exhibit large surface
area and high porosity.18 Considering the diverse structures of pure MOFs, high
tunability on structures and compositions of MOF derivatives allow the creation of
various nanomaterials, which possess heterogeneous sensing structures or new
catalytic function.

PURE MOFS FOR CHEMIRESISTIVE SENSORS


3D MOFs
MOFs have record-breaking large surface areas (up to 8,000 m2/g)7; thus, they
have attracted much attention in broad applications particularly relying on the sur-
face reactions.35 In this regard, MOFs are emerging as next-generation gas-sensing
materials. The chemiresistive sensing properties of pure MOFs were first reported by
Chen et al.36 They synthesized a Co-based zeolite imidazole framework (ZIF-67) that
consists of Co ions and methylimidazole linkers and used it as a formaldehyde sensor
(Figure 2A). Note that low ppm levels of formaldehyde can cause sick building syn-
drome.37 To transduce the sensing signals, the sensors were operated at 150 C
because ZIF-67 is not conductive at room temperature because of its electronic
band gap (1.98 eV) and the poor overlap of electron orbitals.38 Importantly, the
sensors showed the detection limit of 5 ppm with the noticeable response of 1.8 (Fig-
ure 2B), which was attributed to the large surface area (1,800 m2/g) of ZIF-67. In
addition, the sensing performance was independent to relative humidity (RH) up
to 70% RH. The same group further investigated the sensing properties of cobalt-
imidazole frameworks (Co[(im)2]n) synthesized by the assembly of imidazole and co-
balt (II) acetate (Figure 2C).39 The Co[(im)2]n exhibited selective sensing properties

Chem 5, 1–26, August 8, 2019 5


Please cite this article in press as: Koo et al., Metal-Organic Frameworks for Chemiresistive Sensors, Chem (2019), https://doi.org/10.1016/
j.chempr.2019.04.013

Figure 2. 3D MOF-Based Sensors and Their Sensing Properties


(A) Crystal structure of ZIF-67.
(B) Dynamic resistance changes of ZIF-67 to 5‒500 ppm of formaldehyde at 150  C.
(C) Crystal structure of [Co(IM) 2 ] n .
(D) Response versus trimethylamine concentration (inset is response to 2‒50 ppm) at 75  C.
(E) Schematic illustration of NH 2 -UiO-66.
(F) Gas-sensing characteristics of NH 2 -UiO-66 for 10 ppm SO 2 , NO2 , and CO 2 at 150  C.
Reprinted with permission from Chen et al. 36 Copyright 2014 American Chemical Society, Chen et al., 39 Copyright 2014 American Chemical Society, and
DMello et al. 40 Copyright 2019 Royal Society of Chemistry.

toward trimethylamine (Rgas/Rair = 2 to 2 ppm) at an operating temperature of 75 C


(Figure 2D). The sensors exhibited a stable response in various humidity ranges,
similar to ZIF-67-based formaldehyde sensors. Amine functionalized Zr-based
MOFs [Zr6(O)4(OH)4(1,4-benzenedicarboxylate-NH2)6, NH2-UiO-66] also exhibited
chemiresistive sensing characteristics to sulfur dioxide (SO2) at 150 C in Ar atmo-
sphere (Figure 2E).40 The high acidity of SO2 enables a charge transfer coupling
with amine groups in the organic ligands of MOFs, thus leading to the resistance
decrease (jDR/R0j = 21.6% to 10 ppm of SO2) of NH2-UiO-66 (band gap =
2.75 eV)41 upon SO2 adsorption (Figure 2F). Although these works demonstrated
the feasibility of MOF-based chemiresistive sensors, there were some limitations:
(1) the origin of selectivity remains unclear, (2) the need for a heating system compli-
cates the sensing devices, and (3) much enhanced sensitivity is needed to detect
sub-ppm levels of analytes.

In addition, there are bigger challenges in pure MOFs for applications in chemiresis-
tive sensors: poor electrical conductivity and chemical stability. Most MOFs are not
conductive at ambient atmospheres because of the hard metal ions in MOFs and the
poor orbital overlap that restricts the facile transport or flow of electrons.42 There-
fore, the ultrahigh surface area and high porosity of MOFs are not fully involved in
the electrical signal transductions for chemiresistive detection of gas molecules. In
addition, the poor chemical stability of most MOFs hinders practical applications
in MOF-based sensors under harsh conditions.7 Although the integration of MOFs
with conductive materials having high chemical stability enabled the detection of
NH3 gas at room temperature in air,43 they exhibited poor sensing properties.
Therefore, to address the inherent issues in pure MOFs, highly conductive MOFs
with high chemical stability and MOF derivatives, including carbon composites

6 Chem 5, 1–26, August 8, 2019


Please cite this article in press as: Koo et al., Metal-Organic Frameworks for Chemiresistive Sensors, Chem (2019), https://doi.org/10.1016/
j.chempr.2019.04.013

Figure 3. 2D Conductive MOFs and Their Sensing Properties


(A) Crystal structure of M 3 (HHTP) 2 and M 3 (HITP)2 (M = Cu or Ni).
(B) Pattern recognition of diverse analytes using 2D conductive, MOF-based sensing data.
(C) Sensing performance of 2D conductive MOFs toward 16 different VOCs.
Reprinted with permission from Campbell et al.47 Copyright 2015 American Chemical Society.

and metal oxides, have been developed for applications in chemiresistors. These
materials are discussed in the following and other sections.

2D Conductive MOFs
Very recently, MOFs with high electrical conductivity have been reported in a num-
ber of articles.44–46 In general, there are two methods for increasing the electrical
conductivity of MOFs: (1) through-bond and (2) through-space.42 The through-
bond method involves increasing the overlap of the electron orbitals of metal nodes
and organic linkers, whereas the through-space method involves generating an elec-
trical pathway (non-covalent interactions) by adding electroactive fragments in the
free space of MOFs. These approaches can significantly improve the electrical con-
ductivity of MOFs, enabling broad applications of the conductive MOFs in diverse
electronic devices. The Dinca group first reported 2D conductive MOF-based chem-
iresistive sensor, which was capable of detecting ammonia at room temperature.24
The 2D conductive MOFs have higher electrical conductivity compared with pure
MOFs because of charge delocalization and extended p-conjugation between metal
nodes and ligands. As-synthesized 2D Cu3(HITP)2 (HITP = 2,3,6,7,10,11-hexaimino-
triphenylene) exhibited an electrical conductivity of 0.2 S/cm. The 2D Cu3(HITP)2 de-
tected 0.5 to 10 ppm of ammonia with a noticeable response (DG/G0 = 0.1% to
2.5%). In addition, they developed sensor arrays that consisted of 2D Cu3(HHTP)2
(HHTP = 2,3,6,7,10,11-hexahydroxytriphenylene), 2D Cu3(HITP)2, and Ni3(HITP)2
(Figure 3A).47 Upon exposure to 16 volatile organic compounds (VOCs), the sensors
successfully recognized and classified the VOCs into 5 categories (alcohols, ketones
and/or ethers, amines, aromatics, and aliphatics) by using the principal-component
analysis (PCA) (Figure 3B). These results demonstrate that the conductive MOFs are
a new class of materials for chemiresistive sensing. The sensing mechanisms of 2D

Chem 5, 1–26, August 8, 2019 7


Please cite this article in press as: Koo et al., Metal-Organic Frameworks for Chemiresistive Sensors, Chem (2019), https://doi.org/10.1016/
j.chempr.2019.04.013

conductive MOFs are mainly attributed to the charge transfer between analytes and
metal sites, which is demonstrated by the opposite conductance change of Cu- and
Ni-based conductive MOFs (d8 NiII versus d9 CuII) (Figure 3C). In addition, the con-
centration-dependent sensing behavior was also observed: Cu3(HITP)2 exhibited an
increase in conductance toward 200 ppm of n-butylamine, while it exhibited a
decrease in conductance toward 2500 ppm of n-butylamine. The authors assumed
that the hydrogen bonding between the organic ligands and the target molecules
contributed to the sensing properties. Moreover, Rubio-Gimenez et al.26 recently
reported that the chemiresistive behavior of 2D conductive MOFs is strongly linked
to the interaction of Cu(II) sites and gas molecules. They calculated the optimum
geometry and electronic density of states of conductive MOFs upon the adsorption
of specific molecules. The calculation showed that NH3 and H2O can coordinate with
Cu(II) open coordination sites, causing the expansion of a unit cell along the c axis,
and the decrease of conductance of 2D MOFs with the increased electronic band
gap. Overall, the sensing mechanisms of 2D conductive MOFs are unique,
compared with other conventional sensing materials.

However, 2D conductive MOFs synthesized by solution-based reactions still have


some disadvantages such as a low response, insufficient stability, and poor repro-
ducibility. This is mainly originated from the difficulty in the integration of 2D
conductive MOFs with sensing devices without the loss of activity or conductivity.
To facilitate the fabrication process of 2D conductive MOF-based chemiresistors,
the Mirica group reported the direct self-assembly of Cu3(HHTP)2 on shrinkable
polymer films.48 The graphite electrodes were easily drawn on the polymers by using
4B pencils, and Cu3(HHTP)2 or Ni3(HHTP)2 was self-assembled on the polymers to
fabricate conductive MOF-based chemiresistors. The sensors showed a noticeable
response to 80 ppm of NH3 (DI/I0 = 0.7% for Cu3[HHTP]2), NO (jDI/I0j = 1.8% for
Cu3[HHTP]2), and H2S (DI/I0 = 4.2% for Ni3[HHTP]2). Mirica and co-workers further
demonstrated the facile synthesis of conductive MOFs on textiles.51 2D conductive
MOFs were grown on woven cotton fabrics by the self-assembly of metal nodes (Ni)
and organic ligands (HHTP or HITP) in solution (Figure 4A). Because the 2D conduc-
tive MOFs were tightly immobilized on the fabrics, the composite textiles showed
reliable conductance, high flexibility, and good mechanical stability even to
washing. In addition, they showed a high response to 80 ppm of NO (DG/G0 =
81% for Ni3[HITP]2) and H2S (DG/G0 = 98% for Ni3[HTTP]2) (Figure 4B). Furthermore,
the sensors exhibited humidity-independent sensing properties up to 5,000 ppm of
water vapor (18% RH). The theoretical LOD of the sensors was 0.16 ppm for NO
and 0.23 ppm for H2S. Meanwhile, Xu et al.49 reported the layer-by-layer assembly
of 2D conductive MOFs by using a spray-coating method. The sequential spray-
coating of metal nodes and organic ligands dissolved solution produced crystalline
Cu3(HHTP)2 nanofilms with a thickness increase of 2 nm per cycle (Figure 4C). The
thickness of Cu3(HHTP)2 was easily controlled in the range of 20–100 nm by modu-
lating the number of spray-coating cycles. The 2D conductive MOFs with a thickness
of 20 nm exhibited high NH3-sensing properties (jDR/R0j = 45% to 10 ppm) with a
low detection limit (0.5 ppm) (Figure 4D) because the electrical transport in thin
2D conductive MOFs was susceptible to NH3 adsorption. In addition, the NH3-
sensing properties with high stability were observed after 3 months.

Table 1 summarizes the sensing properties of pure MOFs, including non-conductive


and conductive MOFs. The pure MOF-based sensors exhibited a high response to
NO, NH3, and H2S. The response values of pure MOFs are comparable to those of
other promising sensing materials operated in air, such as CNTs and graphene
composites. However, in spite of the many efforts to employ pure MOFs as

8 Chem 5, 1–26, August 8, 2019


Please cite this article in press as: Koo et al., Metal-Organic Frameworks for Chemiresistive Sensors, Chem (2019), https://doi.org/10.1016/
j.chempr.2019.04.013

Figure 4. 2D Conductive MOF-Based Sensing Devices


(A) Fabrication process of MOF-assembled cotton fabrics and their photos and SEM images.
(B) Sensing properties of MOF-cotton composites-based sensors.
(C) Schematic illustration of synthesis of Cu 3 (HHTP) 2 thin film gas sensors by using layer-by-layer spray-coating method.
(D) NH 3 -sensing properties of Cu 3 (HHTP) 2 nanofilms.
Reprinted with permission from Smith et al. 51 Copyright 2017 American Chemical Society and Yao et al. 49 Copyright 2017 Wiley-VCH.

chemiresistors, the pure MOF-based sensors still need further improvements: (1) the
MOFs offer gas responses to some specific gas molecules (e.g., NO, NH3, and H2S),
thus selective detections of the pure MOFs toward the other gas molecules are
limited; (2) the response, particularly the LOD, is low compared with other sensing
materials; and (3) the stability and reproducibility should be further improved.
Very recently, a variety of conductive MOFs have been studied and developed25,52;
therefore, there is a lot of room for the in-depth study of MOF-based chemiresistors.
For instance, Kung et al.53 synthesized 3D conductive MOFs by adding conducting
agents in the void space of the 3D MOFs. Although their sensing properties were not
so dramatic compared with 2D conductive MOF-based sensing materials, they
demonstrated the possibility of 3D conductive MOF-based chemiresistive sensing.
In addition, considering the fact that the sensing performance of conventional

Chem 5, 1–26, August 8, 2019 9


Please cite this article in press as: Koo et al., Metal-Organic Frameworks for Chemiresistive Sensors, Chem (2019), https://doi.org/10.1016/
j.chempr.2019.04.013

Table 1. The Summary of the Sensing Properties in Recent Studies on Pure MOFs for Chemiresistive Sensors
Gas Species Materials Balance Gas Operating Temperature Response Detection Reference
Limit
Trimethylamine [Co(imidazole)2]n air 75 C Rgas/Rair = 15 at 100 ppm 2 ppm Chen et al.39

Formaldehyde ZIF-67 air 150 C Rgas/Rair = 15 at 100 ppm 5 ppm Chen et al.36
Sulfur dioxide NH2-UiO-66 N2 150 C jDR/R0j = 22% at 10 ppm 1 ppm DMello et al.40

Ammonia Cu-BTC/graphene air room temperature DRgas/Rair = 4% at 100 ppm Travlou et al.43
100 ppm
Cu3(HITP)2 N2 room temperature DG/G0 = 2.5% at 10 ppm 500 ppb Campbell et al.24
Cu3(HHTP)2 nanofilms air room temperature jDR/R0j = 45% at 10 ppm 500 ppb Yao et al.49
(thickness = 20 nm)
Methanol Cu3(HHTP)2 N2 room temperature DG/G0 = 9% at 200 ppm N/A Campbell et al.47
Ethanol Cu3(HITP)2 N2 room temperature DG/G0 = 4% at 200 ppm N/A Campbell et al.47
Ni3(HITP)2 N2 room temperature DG/G0 = 4% at 200 ppm N/A Campbell et al.47
Nitrogen monoxide Cu3(HHTP)2 nanorods N2 room temperature jDI/I0j= 1.8% at 80 ppm 2.5 ppm Smith et al.48
Co3(HHTP)2/graphite N2 room temperature DG/G0 = 10% at 80 ppm 5 ppm Ko et al.50
Fe3(HHTP)2/graphite N2 room temperature DG/G0 = 10% at 80 ppm 5 ppm Ko et al.50
Ni3(HITP)2 on cotton N2 room temperature DG/G0 = 81% at 80 ppm 160 ppb Smith and
Mirica51
Hydrogen sulfide Ni3(HHTP)2 nanorods N2 room temperature DI/I0 = 4.2% at 80 ppm 2.5 ppm Smith et al.48
Ni3(HHTP)2 on cotton N2 room temperature DG/G0 = 98% at 80 ppm 230 ppb Smith and
Mirica51
Ni3(HITP)2 on cotton N2 room temperature DG/G0 = 97% at 80 ppm 520 ppb Smith and
Mirica51

Cu-BTC is a Cu-based MOF [Cu3(benzene tricarboxylate)2].

sensing materials can be drastically improved by the functionalization of catalysts or


selectors,1,27 we expect that the introduction of rationally designed catalysts or se-
lectors in pure MOFs can provide a new possibility to overcome the aforementioned
challenges.

MOF MEMBRANES FOR SELECTIVE CHEMIRESISTORS


General Design Principle
So far, MOF-based membranes have been widely studied for gas storage and gas
separation54–56 because MOFs with plenty of micropores have unique properties
for gas adsorption and filtration. Gas molecules can be easily adsorbed on a number
of active sites in MOFs, such as metal nodes and functional groups in organic li-
gands. In addition, MOFs can be engineered to enable selective separation of a
target gas molecule from other mixed gases by rationally designing the pore struc-
ture of MOFs. In other words, since the gas diffusion in a pore is greatly influenced by
chemical and physical parameters, such as gas adsorption, kinetic diameters of
gases, and the pore sizes of MOFs,56 the diffusion of gas molecules with different
kinetic diameters can be effectively controlled by tuning the porosity and chemical
properties of MOFs. Thus, the introduction of MOF membranes can be considered
as the most ideal solution to solve the inherent issue associated with poor selectivity
of chemiresistive sensors. We summarize two types of MOF-based membranes for
development of selective chemiresistors, i.e., MOF-metal and MOF-metal oxides
composites.

MOF-Metal Composites
Pd is a well-known metallic resistor for hydrogen (H2) detection.57 The sensing mech-
anisms and remaining challenges of Pd-based H2 sensors are as follows. In brief, the

10 Chem 5, 1–26, August 8, 2019


Please cite this article in press as: Koo et al., Metal-Organic Frameworks for Chemiresistive Sensors, Chem (2019), https://doi.org/10.1016/
j.chempr.2019.04.013

reaction of H2 with Pd generates palladium hydride (PdHx), changing the resistance


of Pd-based sensors.58 The adsorbed hydrogen on the surface of Pd diffuses into the
interstitial sites of Pd crystals [alpha-phase PdHx (a-PdHx) up to 1% of H2] and inter-
rupts electron movements in Pd. The further reaction of H2 with Pd generates the
beta-phase PdHx (b-PdHx, over 2% of H2 ) with huge volume expansion, where
hydrogen diffuses into the substitutional sites of Pd. By using the volume expansion
of b-PdHx, Pd-based sensors with disconnected Pd junctions were able to detect H2
(over 2%) with ultrafast speed (<75 ms).59 On the other hand, at low levels of H2
(below 1%), the resistance changes of a-PdHx depend on the adsorption of H2. How-
ever, oxygen molecules (O2) in air can be adsorbed on the active sites of Pd, blocking
the reaction sites and resulting in sluggish H2 adsorption.58,60 Therefore, it is crucial
to minimize the negative effect of oxygen on Pd-based sensors. In this regard, Koo
et al.61 demonstrated the use of ZIF-8 as a molecular sieving layer to allow selective
penetration of H2 to Pd-based sensors. They synthesized Pd nanowires by using lith-
ographically patterned nanowire electrodeposition,62 and, thereafter, ZIF-8 was
directly grown on the Pd nanowires (Pd NWs@ZIF-8) (Figures 5A and 5B). Since
ZIF-8 has a pore size of 0.34 nm, the diffusion of H2 (with a kinetic diameter of
0.289 nm) in ZIF-8 membrane is much faster than that of O2 (with a kinetic diameter
of 0.346 nm), enabling the molecular sieving effect (Figure 5C). Therefore, the nega-
tive effect of oxygen on H2-sensing of Pd NWs is remarkably reduced by using the
ZIF-8. Although the response of Pd NWs@ZIF-8 decreased slightly due to the growth
of ZIF-8 on the surface of Pd NWs (Figure 5D), the H2-sensing speed of Pd NWs was
dramatically accelerated after the ZIF-8 deposition; the response time for 1% of H2
was reduced from 164 s to 7 s (Figure 5E), and the recovery time for 1% of H2 was also
decreased from 229 to 10 s (Figure 5F).

MOF-Metal Oxide Composites


As a family of chemiresistive gas-sensing materials, metal oxides, including ZnO,
SnO2, WO3, and In2O3, have been explosively developed for numerous industrial
and domestic applications owing to their relatively high sensitivity, low cost, fast
response, simple operation principle, and real-time operability.63 However, low
selectivity and vulnerability to humidity are major challenges for the practical appli-
cation of metal oxide-based gas sensors. To address these issues, the introduction
of MOF membranes on metal oxides have been considered as a promising approach
for enhancing the selectivity. Yao et al.64 developed ZIF-CoZn loaded ZnO nano-
wires (ZnO@ZIF-CoZn) as selective acetone detecting layers with high stability to
water vapors (Figure 12A). The thickness of ZIF overlayer grown on ZnO nanowire
was controlled in the range of 5–100 nm by changing the concentration of 2-meth-
ylimadazole. Due to the hydrophobic nature of ZIF-CoZn, the ZIF-CoZn membrane
can serve as a filtration layer against water molecules, thus showing humidity-inde-
pendent acetone-sensing characteristics at 260 C (Figure 6B). Furthermore, Zn ions
and Co ions in ZIF-CoZn can act as catalysts for the dissociation and activation of
oxygen molecules, which may lead to significantly enhanced gas-sensing perfor-
mance of ZnO nanowires. Meanwhile, Tian et al.65 reported ZIF-8 coated ZnO
nanowires as a selective formaldehyde detector (Figure 6C). The formaldehyde
gas molecules, which have a kinetic diameter of 0.243 nm, easily pass through the
ZIF-8 layers. On the other hand, other VOCs with larger kinetic diameters, such as
ethanol (0.453 nm), methanol (0.363 nm), acetone (0.46 nm), and toluene
(0.525 nm), can be effectively screened by the ZIF-8 membrane. Therefore, ZIF-8-
coated ZnO nanowires exhibited high selectivity to formaldehyde (Figure 6D) at
300 C, while pristine ZnO nanowires showed a high response to ethanol (Figure 6E).
Although ammonia molecules (with a kinetic diameter of 0.29 nm) can also diffuse
through the ZIF-8 layer, the low activity of ZnO nanowires to ammonia resulted in

Chem 5, 1–26, August 8, 2019 11


Please cite this article in press as: Koo et al., Metal-Organic Frameworks for Chemiresistive Sensors, Chem (2019), https://doi.org/10.1016/
j.chempr.2019.04.013

Figure 5. MOF-Pd Composites and Corresponding H2-Sensing Performance


(A) Schematic illustration of the synthesis of Pd NWs covered by ZIF-8.
(B) SEM image of Pd NWs@ZIF-8.
(C) Schematic illustration of the sensing model of the Pd NWs with and without ZIF-8 layer.
(D–F) H2 -sensing properties of the sensors at room temperature in air: (D) response, (E) response time, and (F) recovery time.
Reprinted with permission from Koo et al. 61 Copyright 2017 American Chemical Society.

the high selectivity toward formaldehyde. In addition, DMello et al.66 synthesized


SnO2 nanoparticle-encapsulated ZIF-67 (SnO2@ZIF-67) to utilize the selective CO2
sorption properties of ZIF-67. The precipitation reaction of Co precursors and
2-methlyimidazole with SnO2 nanoparticles (6.5 nm) resulted in the formation of
SnO2@ZIF-67. The SnO2@ZIF-67 exhibited a CO2 response approximately 2-fold
higher (DR/R0 = 16.5% to 0.5% of CO2 at 205 C) than that of pristine SnO2 nanopar-
ticles (jDR/R0j = 8.8%), although the selectivity of the sensors was not discussed in
the paper. Since the abundant CO2 adsorption sites of ZIF-67 offer catalytic effect,
enhanced response at 205 C is attributed to the higher CO2 sorption capability of
SnO2@ZIF-67 (12.7 cm2/g at 25 C) than SnO2 nanoparticles (1.48 cm2/g).

Based on these results, one can expect that the introduction of MOF overlayer on
metal- or metal-oxide-based sensing layers is a powerful approach to address the

12 Chem 5, 1–26, August 8, 2019


Please cite this article in press as: Koo et al., Metal-Organic Frameworks for Chemiresistive Sensors, Chem (2019), https://doi.org/10.1016/
j.chempr.2019.04.013

Figure 6. MOF-Metal Oxide Composites and Corresponding Gas-Sensing Performance


(A) Schematic illustration for synthesis of ZIF-CoZn coated ZnO nanowires.
(B) Acetone-sensing properties of ZIF-CoZn coated ZnO at 260  C.
(C) Schematic illustration of the molecular sieving effect of ZIF-8-coated ZnO nanowires.
(D) Selectivity property of ZIF-8 coated ZnO nanowires at 300  C.
(E) Selectivity property of ZnO nanowires at 300  C.
Reprinted with permission from Yao et al. 64 Copyright 2016 Wiley-VCH and Tian et al. 65 Copyright 2016 American Chemical Society.

poor selectivity issue of chemical sensors. So far, a few sensing materials, which inte-
grate with MOF membranes for chemiresistive sensors, have been synthesized
(Table 2). Considering the fact that the gas storage and separation of MOFs have
been strongly established and studied,8,70 there is a great possibility to solve critical
challenges in chemical sensors if diverse MOFs can be uniformly deposited on
various sensing materials.

MOF DERIVATIVES FOR CHEMIRESISTIVE SENSORS


Carbon Composites
Porous carbon composites derived from the pyrolysis of MOFs have attracted enor-
mous attention owing to their great chemical and thermal stability, large surface
area, and high porosity.71,72 In particular, since active functional materials can be
easily encapsulated in numerous cavities of the MOFs,73 the desired active mate-
rials, such as metal nanoparticles and TMDs, can be simply loaded on MOF-derived
carbon materials after simple pyrolysis. So far, the active materials-loaded MOF-
derived carbon composites have been widely investigated for improving the mate-
rial’s characteristics for various applications, including electrocatalysts, energy stor-
age systems, and chemical sensors.18 Koo et al.74 reported 2D few-layered WS2
confined in porous carbon composites by using ZIF-67 as sacrificial templates.

Chem 5, 1–26, August 8, 2019 13


Please cite this article in press as: Koo et al., Metal-Organic Frameworks for Chemiresistive Sensors, Chem (2019), https://doi.org/10.1016/
j.chempr.2019.04.013

Table 2. The Summary of the Sensing Properties Reported in Recent Literature of MOF Membranes for Chemiresistive Sensors
Gas Species Materials Strategy Balance Gas Operating Sensing Property Reference
Temperature
Acetone ZnO@ZIF-CoZn (T = 5 nm) durability to air 250 C DR/Rgas = 27.5 at 10 ppm Yao et al.64
humidity independent to 0%–90% RH
Formaldehyde ZnO@ZIF-8 nanowire selectivity air 300 C Rgas/Rair = 12.5 at 100 ppm Tian et al.65
Sratio: improved from 0.74 to 2
Carbon SnO2@ZIF-67 enhanced air 205 C DR/R0 = 16.5% at 0.5% DMello
dioxide adsorption 2-fold higher than pristine SnO2 et al.66
Hydrogen Pd nanowires@ZIF-8 speed air room tres = 7 s & trec = 8 s at 1% Koo et al.61
(T = 160 nm) temperature 20-fold faster than pristine Pd
ZnO@ZIF-8 nanowire selectivity air 300 C Rgas/Rair = 1.44 at 50 ppm no Drobek
(T = 150 nm) response to some gases et al.67
ZnO@ZIF-8 nanorods selectivity air 250 C DI/I0 = 80% at 50 ppm Zhou
(T =  85 nm) Sratio: improved from 0.45 to 2.0 et al.68
ZnO@Pd@ZIF-8 nanowires selectivity air 200 C Rgas/Rair = 6.7 at 50 ppm no response Weber
to some gases et al.69
Ethanol ZnO@ZIF-71 nanorods selectivity air 250 C DI/I0 = 320% at 50 ppm Zhou
(T =  50 nm) Sratio: improved from 1.22 to 1.33 et al.68

T is the thickness of the MOF membrane; Sratio is the selectivity ratio defined by the ratio of the target gas response (Rtarget) to the highest response (Rinterfering)
toward an interfering gas; tres is the response time; trec is the recovery time.

Because of the high activity of WS2, few-layered WS2-loaded carbon composites


exhibit significantly enhanced NO2-sensing properties at room temperature. Since
the WS2 precursors were assembled in the cavity of ZIF-67, the growth of WS2 during
the pyrolysis was effectively suppressed, thereby creating few-layered WS2 function-
alized MOF-derived carbon composites (Figure 7A). As shown in Figure 7B, the
ZIF-67 derived carbon composite materials showed polyhedral morphologies with
a size of 200 nm. The high-resolution transmission electron microscopy (HRTEM)
analysis clearly revealed the single and/or few-layered 2D WS2 on Co, N-doped car-
bon composite materials (Figures 7C and 7D). The WS2 nanoplates-loaded carbon
composites exhibited high NO2 response (DR/Rgas > 48.2% to 5 ppm of NO2) with
excellent selectivity (Figures 7E and 7F) and high stability toward repeated exposure
to NO2. Since the edge site of WS2 has much lower NO2 adsorption energy (1.4 eV)
than that (0.4 eV) of the basal plane of WS2,75 the abundant WS2 edge
sites confined in MOF-derived carbon composites offer significantly promoted
NO2-sensing performances with 10-fold higher reaction kinetics than the pristine
ZIF-67-derived carbon composites.

Metal Oxides
MOF-derived porous metal oxide architectures, which are prepared by the calcina-
tion of MOFs, have attracted great attention in various research fields, including
catalysts,76 energy storage,71 and gas sensors.77 Since MOF-derived metal oxides
manifest mass productivity, high porosity, large surface areas, controllable compo-
sitions, and tunable pore sizes, they can be considered as one of the most promising
gas-sensing materials. Lu et al.78 first reported MOF-templated metal oxides for
chemiresistive sensors. They fabricated porous and hollow Co3O4 concave nano-
cubes by calcining the ZIF-67 templates in air (300 C for 3 h) (Figure 8A). The ob-
tained Co3O4 particles exhibited a concave nanocube structure with a large surface
area (Figure 8B), similar to the polyhedron structure of ZIF-67. Because of the highly
porous structure and high surface area (120.9 cm2/g) of MOFs derived Co3O4
concave nanocubes, highly selective ethanol-sensing characteristics were observed
at 300 C in air (Figure 8C). They also found that the calcination temperature of MOFs

14 Chem 5, 1–26, August 8, 2019


Please cite this article in press as: Koo et al., Metal-Organic Frameworks for Chemiresistive Sensors, Chem (2019), https://doi.org/10.1016/
j.chempr.2019.04.013

Figure 7. MOF-Derived Carbon Composites and Corresponding Sensing Performances


(A) Schematic illustration of the synthesis of MOF-derived carbon/WS 2 composites.
(B) TEM image, (C) magnified TEM image, and (D) HRTEM image of MOF-derived carbon composites functionalized with WS 2 nanoplates.
(E) Dynamic response transitions of the sensors toward 1–5 ppm of NO2 at room temperature.
(F) Response of the sensors to various analytes.
Reprinted with permission from Koo et al. 74 Copyright 2018 Wiley-VCH.

plays an important role in optimizing surface area and porosity of MOF-derived


metal oxides. The crystallinity of the MOF-derived oxides increased as the calcina-
tion temperature increased, but the surface area and porosity of the oxides
decreased. As a result, the Co3O4 nanocubes calcined at 300 C showed higher
ethanol response (Rair/Rgas = 1.8 to 10 ppm) than other samples calcined at 400 C
and 500 C (Figure 8D). Moreover, the composition of MOF-derived metal oxides
can be controlled by simply changing metal ion species consisting of MOFs. For
example, the calcination of MIL-68, MIL-88A, and MOF-5, which contain In, Fe,
and Zn ions, respectively, results in the formation of highly porous In2O3, ZnO,
and Fe2O3.79–81 Selective detections of formaldehyde using In2O3 and acetone us-
ing ZnO and Fe3O4 were investigated.

To further enhance the sensing properties of MOF-derived metal oxides, Jang


et al.82 reported ZnO-Co3O4 hierarchical structures obtained from the hybridization
of heterogeneous MOFs. The precipitation reaction of the as-synthesized ZIF-67
rods, Zn ions, and 2-methylimidazole produced hierarchical MOFs with Zn and Co
nodes (Figure 9A). The continuous growth of Zn-based, leaf-like ZIF (ZIF-L) in the
center of Co-based ZIF-67 rods generated hierarchical and heterogeneous MOFs
(Figure 9B). Subsequently, the calcination of hierarchical MOFs induced the hierar-
chical structures consisting of n-type ZnO sheets and p-type Co3O4 rods (Figure 9C).
The MOF-derived ZnO-Co3O4 exhibited enhanced acetone-sensing performance
(Rair/Rgas = 29.0 to 5 ppm) compared to that (Rair/Rgas = 1.05) of pure ZIF-67 rods-
derived metal oxides (Co3O4) measured at 450 C (Figure 9D) because n-type ZnO

Chem 5, 1–26, August 8, 2019 15


Please cite this article in press as: Koo et al., Metal-Organic Frameworks for Chemiresistive Sensors, Chem (2019), https://doi.org/10.1016/
j.chempr.2019.04.013

Figure 8. MOF-Derived Metal Oxides and Corresponding Sensing Performance


(A) Schematic illustration for the synthesis of ZIF-67 derived Co 3 O 4 concave nanocubes.
(B) SEM image of ZIF-67 derived Co 3 O4 concave nanocubes.
(C) Response of Co 3 O 4 concave nanocubes to four gas species at 300  C.
(D) Response of Co 3 O 4 concave nanocubes to various concentrations of ethanol at 300  C.
Reprinted with permission from Lu et al. 78 Copyright 2014 American Chemical Society.

and p-type Co3O4 create additional electron depletion layers (n-p junction) that can
cause large resistance changes upon exposure to target analytes.83 Despite the
effective utilization of the large surface area and high porosity of MOF-derived metal
oxides, MOF-derived metal oxides still exhibit unsatisfactory sensing properties.
This may be attributed to the low activity of the surface reactions between MOF-
derived metal oxides and analytes. Thus, the use of catalysts is needed to promote
the surface reactions.

Because MOFs have high porosity and numerous cavities in their structures, catalytic
metal nanoparticles can be easily encapsulated in MOFs.84 The ultrasmall and well-
dispersed metal nanoparticles in MOFs (metal@MOFs) exhibit unexpected proper-
ties in diverse applications. In general, there are four representative methods for the
encapsulation of metal nanoparticles in MOFs: (1) ship-in-bottle, (2) bottle-around-
ship, (3) sandwich assembly, and (4) in-situ encapsulation approaches (Figure 10A).85
Briefly, the ship-in-bottle strategy is based on the assembly of target metal nanopar-
ticles in a MOF matrix that has already been formed. The bottle-around-ship strategy
is the assembly of MOF around metal nanoparticles. The sandwich assembly is the
decoration of metal nanoparticles on a MOF (core) and the subsequent overgrowth
of a MOF (shell). Last, the in-situ encapsulation route involves mixing all precursors
and simultaneous growth of MOF and metal nanoparticles. To use metal@MOFs in
the applications in chemiresistive sensors, Koo et al.86 first reported metal nanopar-
ticles-doped metal oxides for acetone sensors by using Pd nanoparticles-embedded
ZIF-67 (Pd@ZIF-67) as a sacrificial template. Because of the small size (11.6 Å) of the
cavities in ZIF-67,84 the growth of Pd nanoparticles in ZIF-67 was restricted; thus,
ultrasmall (2–3 nm) Pd nanoparticles with high uniformity were produced (Figures

16 Chem 5, 1–26, August 8, 2019


Please cite this article in press as: Koo et al., Metal-Organic Frameworks for Chemiresistive Sensors, Chem (2019), https://doi.org/10.1016/
j.chempr.2019.04.013

Figure 9. Hierarchical and Heterogeneous MOF-Derived Metal Oxides and Corresponding Sensing Performance
(A) Schematic illustration of the synthesis of hierarchical MOFs and their derivatives.
(B) TEM image of hierarchical MOFs.
(C) TEM image of MOF-derived ZnO-Co 3 O 4 hierarchical structures.
(D) Acetone-sensing properties of MOF-derived ZnO-Co 3 O 4 hierarchical structures at 450  C.
Reprinted with permission from Jang et al.82 Copyright 2018 American Chemical Society.

10B and 10C). After encapsulation of the Pd nanoparticles in ZIF-67, the direct
calcination of Pd@ZIF-67 was conducted at 400 C for 1 h, resulting in the formation
of PdO nanoparticles functionalized Co3O4 hollow nanocubes (Figures 10D and
10E). Owing to the catalytic effect of PdO nanoparticles, the PdO loaded Co3O4 hol-
low nanocubes showed enhanced acetone-sensing characteristics at 350 C, in terms
of response (Rair/Rgas = 2.51 at 5 ppm) and selectivity (Figures 10F and 10G). The
hollow structure of Co3O4 facilitated the diffusion of gas molecules, whereas the
PdO nanoparticles promoted the surface reaction of the active sensing layers toward
acetone molecules (Figure 10H). In particular, PdO nanoparticles, as well-known
electronic sensitizers,87 deprive electrons of Co3O4 when exposed to air and donate
electrons to Co3O4 when exposed to acetone (reducing gas), hence effectively
improving the sensing properties through the catalytic effect of the PdO
nanoparticles.

In addition, the sensing performances of MOF-derived catalytic metal nanoparticles-


loaded metal oxides can be further improved by developing heterogeneous
structures. It is noted that most MOF-derived metal oxides possess p-type semicon-
ducting characteristics, which show poor sensing performances compared with
n-type SMOs.83,88 Therefore, it is beneficial to combine the MOF-derived p-type
metal oxides with SnO2 or WO3, which are well-known metal oxides for

Chem 5, 1–26, August 8, 2019 17


Please cite this article in press as: Koo et al., Metal-Organic Frameworks for Chemiresistive Sensors, Chem (2019), https://doi.org/10.1016/
j.chempr.2019.04.013

Figure 10. Catalytic Metal Nanoparticles-Loaded MOFs and Their Derivatives for Chemiresistive Sensors
(A) Schematic illustrations for the mechanisms of the encapsulation of metal nanoparticles in MOFs: (i) ship-in-bottle, (ii) bottle-around-ship, (iii)
sandwich assembly, and (iv) in-situ encapsulation methods.
(B) TEM image and (C) HRTEM image of Pd@ZIF-67.
(D) TEM image and (E) HRTEM image of Pd@ZIF-67 derived PdO loaded Co3 O4 hollow nanocubes.
(F) Acetone-sensing characteristics of PdO loaded Co 3 O4 hollow nanocubes.
(G) Response of PdO loaded Co 3 O 4 hollow nanocubes various analytes at 350  C.
(H) Schematic illustration of the sensing mechanism of PdO loaded Co 3 O 4 hollow nanocubes.
Reprinted with permission from Chen et al.85 Copyright 2017 Royal Society of Chemistry, and Koo et al. 86 Copyright 2017 American Chemical Society.

chemiresistive sensing. To address these issues, the direct surface modification of


MOF-derived metal oxides was employed by using a galvanic replacement reaction.
Jang et al.89 prepared PdO nanoparticles-loaded Co3O4 nanocubes by using
Pd@ZIF-67 as a sacrificial template. Then, the subsequent galvanic replacement
reaction of PdO-loaded Co3O4 nanocubes with Sn-based precursor-dissolved
solutions produced PdO-SnO2-Co3O4 heterogeneous structures (Figures 11A–
11D) because of the potential difference of Co3+/Co2+ (1.87 V) and Sn4+/Sn2+
(0.09 V).90 Interestingly, the n-type PdO-SnO2-Co3O4 heterogeneous structures
showed markedly enhanced acetone-sensing performance (Rair/Rgas = 22.8 at
5 ppm at 450 C) compared with PdO loaded Co3O4 nanocubes (Rair/Rgas = 1.05)
(Figure 11E). Moreover, these structures successfully distinguished acetone mole-
cules from 7 other interfering gases (Figure 11F), demonstrating the high acetone
selectivity of n-type PdO-SnO2-Co3O4.

18 Chem 5, 1–26, August 8, 2019


Please cite this article in press as: Koo et al., Metal-Organic Frameworks for Chemiresistive Sensors, Chem (2019), https://doi.org/10.1016/
j.chempr.2019.04.013

Figure 11. MOF-Based Heterogeneous Structures and Corresponding Sensing Properties


(A) Schematic illustration for the synthesis of PdO loaded SnO 2 -Co 3 O 4 hollow cubes.
(B) SEM image of Pd loaded ZIF-67.
(C) SEM image of PdO loaded Co3 O 4 hollow cubes.
(D) SEM image of PdO loaded SnO 2 -Co 3 O4 hollow cubes.
(E) Acetone-sensing properties of PdO loaded SnO 2 -Co 3 O 4 hollow cubes at 450  C.
(F) Pattern recognition of various analytes by using PCA of the sensing data.
Reprinted with permission from Jang et al.89 Copyright 2017 American Chemical Society.

In addition, Koo et al.91 reported the use of metal@MOFs as a template for the
synthesis of complex catalysts decorated on WO3-based chemiresistors. The met-
al@metal oxides derived from metal@MOFs dramatically promoted the gas-
sensing properties of WO3 nanofibers. In brief, Pd-loaded ZIF-8 (Pd@ZIF-8) was
decorated on electrospun nanofibers consisting of polymers and tungsten (W) pre-
cursors by using an electrospinning method (Figure 12A). As shown in Figures 12B
and 12C, the ultrasmall Pd nanoparticles loaded in ZIF-8 were successfully func-
tionalized on the as-spun nanofibers. The subsequent calcination of Pd@ZIF-8
loaded nanofibers produced WO3 nanofibers functionalized with Pd loaded ZnO
nanocubes (Pd@ZnO) (Figures 12D and 12E). Interestingly, the Pd@ZnO loaded
WO3 nanofibers exhibited superior toluene detection capability (Rair/Rgas =
22.22 to 1 ppm at 350 C) (Figure 12F), with high selectivity and fast response
time (20 s). The large electron depletion region in n-type WO3 nanofibers
was induced by the multi-heterojunction structures of WO3-ZnO, ZnO-Pd, and

Chem 5, 1–26, August 8, 2019 19


Please cite this article in press as: Koo et al., Metal-Organic Frameworks for Chemiresistive Sensors, Chem (2019), https://doi.org/10.1016/
j.chempr.2019.04.013

Figure 12. MOF-Based Heterogeneous Structures and Corresponding Sensing Properties


(A) Schematic illustration for the fabrication of Pd@ZnO loaded WO 3 nanofibers.
(B) TEM image and (C) HRTEM image of Pd@ZIF-8 loaded as-spun nanofiber.
(D) TEM image and (E) HRTEM image of Pd@ZnO loaded WO 3 nanofiber.
(F) Toluene-sensing properties of Pd@ZnO loaded WO 3 nanofibers at 350  C.
(G) Schematic illustration of the sensing mechanism of Pd@ZnO loaded WO 3 nanofibers.
Reprinted with permission from Koo et al. 91 Copyright 2016 American Chemical Society.

WO3-Pd, thus enabling a large resistance variation during the toluene sensing.83 In
addition, Pd nanoparticles were partially oxidized to PdO in air and were easily
reduced to Pd when exposed to toluene, resulting in effective modulation of the
surface depletion layers (Figure 12G).

Table 3 summarizes the gas-sensing properties of MOF-derivatives for chemiresis-


tors reported in the recent literature. Various scientific and technological advance-
ments have been achieved for the development of MOF-derivatives, and the
sensing properties are comparable to the state-of-the-art carbon composites
and metal-oxide-based chemical sensors. However, there are remaining chal-
lenges: high operating temperature, selectivity, and long-term stability. Given
that the researches on MOF-derivatives have received enormous attention in
the last few years, new strategies, such as single-atom catalysts or bimetallic
catalysts,18,99 can shed light on the development of the next-generation
chemiresistors.

20 Chem 5, 1–26, August 8, 2019


Please cite this article in press as: Koo et al., Metal-Organic Frameworks for Chemiresistive Sensors, Chem (2019), https://doi.org/10.1016/
j.chempr.2019.04.013

Table 3. The Summary of the Sensing Properties in the Recent Papers on MOF Derivatives for Chemiresistive Sensors
Gas Species Materials Balance Operating Response Detection Reference
Gas Temperature Limit
Nitrogen ZIF-67 derived carbon composites-loaded N2 room jDR/R0j = 1% at 5 ppm 100 ppb Rui et al.92
dioxide MWCNTs temperature
ZIF-67 derived WS2 functionalized carbon air room DR/Rgas = 18% at 1 ppm 100 ppb Koo et al.74
composites temperature
SWCNTs loaded PdO-Co3O4 derived from air room DR/Rair = 27.3% at 20 ppm 10 ppm Choi
Pd@ZIF-67 temperature et al.93
Ethanol ZIF-67 templated Co3O4 concave air 300 C Rgas/Rair = 3.25 at 200 ppm 10 ppm Lü et al.78
nanocubes
HKUST-1 templated Cu2O/CuO cages air 150 C Rgas/Rair = 6.5 at 200 ppm N/A Wang
et al.94
In/Ga-MOF-68 derived In/Ga oxides air 235 C Rair/Rgas = 80 at 200 ppm 2 ppm Cui et al.95

Benzene MOF-5 templated ZnO nanocages air 300 C DR/Rair = 1.28 at 0.1 ppm 100 ppb Li et al.81
ZIF-8 templated hollow ZnO air 450 C DR/Rgas = 1.9 at 5 ppm N/A Li et al.96

Formaldehyde MOF templated mesoporous In2O3 air 210 C DR/Rgas = 7.5 at 10 ppm N/A Wang
nanorod et al.79
p-Xylene ZIF-67 derived hollow Co3O4 nanocages air 225 C Rgas/Rair = 78.6 at 5 ppm 250 ppb Jo et al.97
Toluene ZIF-67 derived hollow Co3O4 nanocages air 225 C Rgas/Rair = 43.8 at 5 ppm N/A Jo et al.97
Pd@ZIF-8 templated Pd@ZnO-WO3 air 350 C Rair/Rgas = 22.22 at 1 ppm 0.1 ppm Koo et al.91
nanofibers
Acetone Pd@ZIF-ZnCo derived PdO@ZnO/ZnCo2O4 air 250 C DR/Rair = 69% at 5 ppm 400 ppb Koo et al.88
hollow spheres
Pd@ZIF-67 derived PdO@Co3O4 nanocubes air 350 C Rgas/Rair = 2.51 at 5 ppm 100 ppb Koo et al.86
Pd@ZIF-8 templated PdO@ZnO-SnO2 air 400 C Rair/Rgas = 5.06 at 1 ppm 100 ppb Koo et al.98
nanotubes
Pd@ZIF-67 templated Co3O4-PdO loaded air 450 C Rair/Rgas = 22.8 at 5 ppm 5 ppb Jang
SnO2 nanocubes et al.89
Hierarchical MOF derived ZnO-Co3O4 air 450 C Rair/Rgas = 29 at 5 ppm 1 ppm Jang
et al.82

MWCNTs is multi-walled carbon nanotubes. SWCNTs is single-walled carbon nanotubes.

Conclusions and Perspectives


The field of MOF-based chemiresistive sensors has rapidly grown in the past several
years and has been explored in various approaches: (1) pure MOFs, (2) MOF mem-
branes, and (3) MOF derivatives. Pure MOFs are promising sensing materials
because of the tremendously large surface area and porosity that can enhance the
surface reactions with the analytes. In particular, conductive MOFs are able to effi-
ciently transduce the electrical signals from the surface reactions; thus, they are
considered as a new class of materials for chemiresistive sensors. In addition,
MOF membranes promote the selective penetration of target gases, minimizing
the negative effects of interfering analytes because of their unique gas adsorption
and separation properties. Last, as sacrificial templates, MOFs can be converted
to various MOF-derived nanomaterials, including carbon composites and metal ox-
ides, which are well-established sensing materials. Various strategies have been sug-
gested to increase active sites and promote surface reactions with target gases for
MOF-derivative-based chemical sensors. They include the control of the structure,
the morphology, the composition, the addition of active materials, and the use of
nanoscale catalysts encapsulated in the cavity of MOFs.

In this review, we summarized the recent progress of chemiresistive sensors


based on pure MOFs, MOF membranes, and MOF derivatives. The representative

Chem 5, 1–26, August 8, 2019 21


Please cite this article in press as: Koo et al., Metal-Organic Frameworks for Chemiresistive Sensors, Chem (2019), https://doi.org/10.1016/
j.chempr.2019.04.013

Table 4. Representative MOF-Based Materials for Chemiresistive Sensors and the Comparison of Their Characteristics
Classification Sub-classification Example Synthetic Method Advantages Disadvantages
Pure MOFs 3D MOFs (not  synthesis at r.t.  ultrahigh surface area  poor conductivity
conductive at r.t.)  solvothermal  micro- or mesoporous  poor stability
 hydrothermal structure  high operating temp.
 low sensitivity
 low selectivity

ZIF-6735
conductive MOFs  solvothermal  high surface area  low selectivity
(at r.t.)  hydrothermal  micro- or mesoporous  poor stability
 layer-by-layer spray structure  low reproducibility
coating  operation at r.t.  limited strategies for
 in-situ growth on sup-  noticeable sensitivity at sensor fabrication w/o
ports r.t. a loss of activity of
MOFs
Cu3(HITP)224
(through-bond)
 deposition of con-  high surface area  low sensitivity
ducting agents to  micro- or mesoporous  low selectivity
MOFs structure  poor stability
 hydro- or solvothermal  operation at r.t.  low reproducibility

tin oxide_Nu-100053
(through-space)
MOF MOF-metal  in-situ growth by r.t.  micro- or mesoporous  low response
membranes composites synthesis structure  not uniform deposition
 in-situ growth by hydro-  selective gas separa- of diverse MOFs
or solvothermal tion and adsorption  limited strategies for
 high selectivity or fast efficient fabrication of
sensing speed sensing devices
Pd nanowire@ZIF-861
MOF-metal oxide  in-situ growth by r.t.  micro- or mesoporous  not uniform deposi-
composites synthesis structure tion of diverse MOFs
 in-situ growth by hydro-  enhanced sensitivity  limited strategies for
or solvothermal  enhanced selectivity efficient fabrication of
 durability to humidity sensing devices
 high operating temp.
 poor long-term stability
ZnO@ZIF-CoZn64
MOF carbon composites  integration of MOF  high surface area  vulnerability to humid-
derivatives with active materials  micro- or mesoporous ity
 pyrolysis structure  low reproducibility
 operation at low temp.  slow sensing speed
 noticeable sensitivity at  poor long-term stability
r.t.  poisoning of active ma-
carbon-WS2  improved selectivity terials
composites74
metal oxides  morphology & compo-  meso- or macropo-  high operating temp.
sition control of MOFs rous structure  vulnerability to humidi-
 subsequent calcination  enhanced sensitivity ty
 high thermal/chemical  low selectivity
stability  baseline drift during
long-term operation
Co3O4@ZnO sheet82
 post-treatment of met-  meso- or macropo-  high operating temp.
al@MOF-derived rous structure  vulnerability to humidi-
metal oxides  high sensitivity ty
 improved selectivity  baseline drift during
 high thermal and chem- long-term operation
ical stability  catalyst poisoning
PdO-loaded SnO2-
Co3O4 hollow cubes89
(Continued on next page)

22 Chem 5, 1–26, August 8, 2019


Please cite this article in press as: Koo et al., Metal-Organic Frameworks for Chemiresistive Sensors, Chem (2019), https://doi.org/10.1016/
j.chempr.2019.04.013

Table 4. Continued
Classification Sub-classification Example Synthetic Method Advantages Disadvantages
 electrospinning of  meso- or macropo-  high operating temp.
MOFs or metal@- rous structure  vulnerability to humidi-
MOFs  high sensitivity ty
 subsequent calcination  improved selectivity  baseline drift during
 high thermal and long-term operation
chemical stability  catalyst poisoning

PdO@ZnO loaded
SnO2 nanotubes98

r.t., room temperature; temp, temperature; w/o, without.


Reprinted with permission from Chen et al. 35 Copyright 2014 American Chemical Society, Campbell et al.24 Copyright 2015 Wiley-VCH, Kung et al.53 Copyright
2018 American Chemical Society, Koo et al.61 Copyright 2017 American Chemical Society, Yao et al.64 Copyright 2016 Wiley-VCH, Koo et al.74 Copyright 2016
Wiley-VCH, Jang et al.82 Copyright 2018 American Chemical Society, Jang et al.89 Copyright 2017 American Chemical Society, and Koo et al.98 Copyright 2017
American Chemical Society.

MOF-based materials and the correlation between synthetic methods of each mate-
rial and related sensing characteristics are summarized in Table 4. MOFs for chem-
iresistors have been intensively studied in recent years, and several research groups
have achieved innovative advancements. However, further improvements are still
needed to use MOFs as commercial sensors in a wide range of applications. (1)
Pure MOFs should display a higher response with higher stability and reproduc-
ibility, compared with other conventional sensing materials. In addition, the selec-
tivity of pure MOF-based chemiresistors should be extended to various analytes.
(2) In the case of MOF membrane, facile and robust deposition methods should
be developed to provide uniform and homogeneous coverages of diverse MOF
layers on sensing materials while maintaining minimal disturbance of active sites.
Moreover, establishing the interfaces, surfaces, reaction kinetics, and thermody-
namics mechanisms is required for MOF-membrane-assisted chemiresistors. (3)
MOF-derived carbon composites have some challenges, similar to the conventional
carbon-based sensors: low response, poor stability, and low repeatability. On the
other hand, MOF-derived metal oxides require high operating temperature, which
can complicate the sensing system and cause stability or cross-selectivity issues.
Our view is that (1) the use of catalysts or selectors in conductive MOFs can broaden
the detection capability of pure-MOF-based sensors; (2) 3D conductive MOFs,
which have a higher surface area compared with 2D MOFs,100 can further improve
the sensing performances of pure MOFs; (3) the construction of uniform MOF mem-
branes with a controlled pore size on sensing materials can produce sensor arrays
with excellent selectivity to diverse gas molecules, considering the various MOF
deposition methods, such as chemical vapor deposition, which have been reported
recently101,102; and (4) last, MOF-derivatives, such as a catalyst or heterogeneous
backbone, are expected to further enhance their sensing properties by the integra-
tion with highly active materials such as single-atom catalysts and bimetallic cata-
lysts,103,104 with the careful optimization of structure, morphology, and composition.
Since MOFs have a number of fascinating features for applications in gas sensors and
have a lot of spaces that have not yet been explored, MOFs for chemiresistors will
grow rapidly in the future, overcoming the critical challenges of current chemical
sensors. We hope that this review guides the way for the practical applications of
MOF-based chemiresistive sensors.

ACKNOWLEDGMENTS
This work was supported by the Ministry of Trade, Industry & Energy (Korea) under
the Industrial Technology Innovation Program (No. 10070075). This work was also
supported by Wearable Platform Materials Technology Center (WMC) funded by

Chem 5, 1–26, August 8, 2019 23


Please cite this article in press as: Koo et al., Metal-Organic Frameworks for Chemiresistive Sensors, Chem (2019), https://doi.org/10.1016/
j.chempr.2019.04.013

National Research Foundation of Korea (NRF) Grant of the Korean Government


(MSIP) (No. 2016R1A5A1009926). This work was also funded by the Ministry of Sci-
ence, ICT & Future Planning as a Biomedical Treatment Technology Development
Project (2015M3A9D7067418).

AUTHOR CONTRIBUTIONS
Conceptualization, W.-T.K. and I.-D.K.; Investigation, W.-T.K. and J.-S.J.; Visualiza-
tion, W.-T.K. and J.-S.J.; Writing – Original Draft, W.-T.K. and J.-S.J.; Writing –
Review & Editing, W.-T.K., J.-S.J., and I.-D.K.; Supervision, I.-D.K.; Funding Acquisi-
tion, I.-D.K.

REFERENCES AND NOTES


1. Kim, S.J., Choi, S.J., Jang, J.S., Cho, H.J., and chemical sensing and explosive detection. 25. Aubrey, M.L., Wiers, B.M., Andrews, S.C.,
Kim, I.D. (2017). Innovative nanosensor for Chem. Soc. Rev. 43, 5815–5840. Sakurai, T., Reyes-Lillo, S.E., Hamed, S.M., Yu,
disease diagnosis. Acc. Chem. Res. 50, 1587– C.J., Darago, L.E., Mason, J.A., Baeg, J.O.,
1596. 15. Kreno, L.E., Hupp, J.T., and Van Duyne, R.P. et al. (2018). Electron delocalization and
(2010). Metal–organic framework thin film for charge mobility as a function of reduction in a
2. Loutfi, A., Coradeschi, S., Mani, G.K., Shankar, enhanced localized surface plasmon metal–organic framework. Nat. Mater. 17,
P., and Rayappan, J.B.B. (2015). Electronic resonance gas sensing. Anal. Chem. 82, 8042– 625–632.
noses for food quality: a review. J. Food Eng. 8046.
144, 103–111. 26. Rubio-Giménez, V., Almora-Barrios, N.,
16. Lu, G., and Hupp, J.T. (2010). Metal–organic Escorcia-Ariza, G., Galbiati, M., Sessolo, M.,
3. Brown, S.K., Sim, M.R., Abramson, M.J., and
frameworks as sensors: A ZIF-8 based Tatay, S., Martı́-Gastaldo, C., Almora-Barrios,
Gray, C.N. (1994). Concentrations of volatile
FabryPérot device as a selective sensor for N., Escorcia-Ariza, G., Galbiati, M., Sessolo,
organic compounds in indoor air–a review.
chemical vapors and gases. J. Am. Chem. M., Tatay, S., and Martı́-Gastaldo, C. (2018).
Indoor Air 4, 123–134.
Soc. 132, 7832–7833. Origin of the chemiresistive response of
4. Kostiainen, R. (1995). Volatile organic ultrathin films of conductive metal–organic
compounds in the indoor air of normal and 17. Prestipino, C., Regli, L., Vitillo, J.G., Bonino, frameworks. Angew. Chem., Int. Ed. 57,
sick houses. Atoms. Environ. 29, 693–702. F., Damin, A., Lamberti, C., Zecchina, A., 15086–15090.
Solari, P.L., Kongshaug, K.O., and Bordiga, S.
5. Potyrailo, R.A. (2016). Multivariable sensors (2006). Local structure of framework Cu(II) in 27. Schroeder, V., Savagatrup, S., He, M., Lin, S.,
for ubiquitous monitoring of gases in the era HKUST-1 metallorganic framework: and Swager, T.M. (2019). Carbon nanotube
of internet of things and industrial internet. spectroscopic characterization upon chemical sensors. Chem. Rev. 119, 99–663.
Chem. Rev. 116, 11877–11923. activation and interaction with adsorbates.
Chem. Mater. 18, 1337–1346. 28. Liu, Y., Dong, X., and Chen, P. (2012).
6. Kim, I.-D., Rothschild, A., and Tuller, H.L. Biological and chemical sensors based on
(2013). Advances and new directions in gas- 18. Dang, S., Zhu, Q.-L., and Xu, Q. (2017). graphene materials. Chem. Soc. Rev. 41,
sensing devices. Acta Mater. 61, 974–1000. Nanomaterials derived from metal–organic 2283–2307.
frameworks. Nat. Rev. Mater. 3, 17075.
7. Furukawa, H., Cordova, K.E., O’Keeffe, M., 29. Anichini, C., Czepa, W., Pakulski, D.,
and Yaghi, O.M. (2013). The chemistry and 19. Stassen, I., Burtch, N., Talin, A., Falcaro, P., Aliprandi, A., Ciesielski, A., and Samorı̀, P.
applications of metal-organic frameworks. Allendorf, M., and Ameloot, R. (2017). An (2018). Chemical sensing with 2D materials.
Science 341, 1230444. updated roadmap for the integration of Chem. Soc. Rev. 47, 4860–4908.
8. Li, J.R., Kuppler, R.J., and Zhou, H.C. (2009). metal–organic frameworks with electronic
Selective gas adsorption and separation in devices and chemical sensors. Chem. Soc. 30. Franke, M.E., Koplin, T.J., and Simon. (2006).
metal–organic frameworks. Chem. Soc. Rev. Rev. 46, 3185–3241. Metal and metal oxide nanoparticles in
38, 1477–1504. chemiresistors: does the nanoscale matter?
20. Campbell, M.G., and Dinc a, M. (2017). Metal– Small 2, 36–50.
9. Morris, R.E., and Wheatley, P.S. (2008). Gas organic frameworks as active materials in
storage in nanoporous materials. Angew. electronic sensor devices. Sensors 17, 1108. 31. Ammu, S., Dua, V., Agnihotra, S.R., Surwade,
Chem. Int. Ed. Engl. 47, 4966–4981. S.P., Phulgirkar, A., Patel, S., and Manohar,
21. Zhao, W., Peng, J., Wang, W., Liu, S., Zhao, Q., S.K. (2012). Flexible, all-organic chemiresistor
10. Wang, H., Zhu, Q.-L., Zou, R., and Xu, Q. and Huang, W. (2018). Ultrathin two- for detecting chemically aggressive vapors.
(2017). Metal-organic frameworks for energy dimensional metal-organic framework J. Am. Chem. Soc. 134, 4553–4556.
applications. Chem 2, 52–80. nanosheets for functional electronic devices.
Coord. Chem. Rev. 377, 44–63. 32. Torsi, L., Magliulo, M., Manoli, K., and Palazzo,
11. Lee, J., Farha, O.K., Roberts, J., Scheidt, K.A., G. (2013). Organic field-effect transistor
Nguyen, S.T., and Hupp, J.T. (2009). Metal– 22. Chidambaram, A., and Stylianou, K.C. (2018). sensors: a tutorial review. Chem. Soc. Rev. 42,
organic framework materials as catalysts. Electronic metal–organic framework sensors. 8612–8628.
Chem. Soc. Rev. 38, 1450–1459. Inorg. Chem. Front. 5, 979–998.
33. Lee, J.-H. (2009). Gas sensors using
12. Diercks, C.S., Liu, Y., Cordova, K.E., and
23. Fang, X., Zong, B., and Mao, S. (2018). Metal– hierarchical and hollow oxide nanostructures:
Yaghi, O.M. (2018). The role of reticular
organic framework-based sensors for overview. Sens. Actuators B 140, 319–336.
chemistry in the design of CO2 reduction
environmental contaminant sensing.
catalysts. Nat. Mater. 17, 301–307.
Nanomicro Lett. 10, 64. 34. Swager, T.M. (2018). Sensor technologies
13. Kreno, L.E., Leong, K., Farha, O.K., Allendorf, empowered by materials and molecular
M., Van Duyne, R.P., and Hupp, J.T. (2012). 24. Campbell, M.G., Sheberla, D., Liu, S.F., innovations. Angew. Chem. Int. Ed. Engl. 57,
Metal–organic framework materials as Swager, T.M., and Dinc a, M. (2015). Cu3 4248–4257.
chemical sensors. Chem. Rev. 112, 1105–1125. (hexaiminotriphenylene)2: an electrically
conductive 2D metal–organic framework for 35. Czaja, A.U., Trukhan, N., and Müller, U. (2009).
14. Hu, Z., Deibert, B.J., and Li, J. (2014). chemiresistive sensing. Angew. Chem. Int. Ed. Industrial applications of metal–organic
Luminescent metal–organic frameworks for 54, 4349–4352. frameworks. Chem. Soc. Rev. 38, 1284–1293.

24 Chem 5, 1–26, August 8, 2019


Please cite this article in press as: Koo et al., Metal-Organic Frameworks for Chemiresistive Sensors, Chem (2019), https://doi.org/10.1016/
j.chempr.2019.04.013

36. Chen, E.X., Yang, H., and Zhang, J. (2014). shrinkable polymer films. Chem. Mater. 28, 61. Koo, W.T., Qiao, S., Ogata, A.F., Jha, G.,
Zeolitic imidazolate framework as 5264–5268. Jang, J.S., Chen, V.T., Kim, I.D., and Penner,
formaldehyde gas sensor. Inorg. Chem. 53, R.M. (2017). Accelerating palladium nanowire
5411–5413. 49. Yao, M.S., Lv, X.J., Fu, Z.H., Li, W.H., Deng, H2 sensors using engineered nanofiltration.
W.H., Wu, G.D., and Xu, G. (2017). Layer-by- ACS Nano 11, 9276–9285.
37. Wargocki, P., Wyon, D.P., Baik, Y.K., Clausen, layer assembled conductive metal–organic
G., and Fanger, P.O. (1999). Perceived air framework nanofilms for room-temperature 62. Menke, E.J., Thompson, M.A., Xiang, C.,
quality, sick building syndrome (SBS) chemiresistive sensing. Angew. Chem. Int. Ed. Yang, L.C., and Penner, R.M. (2006).
symptoms and productivity in an office with Engl. 56, 16510–16514. Lithographically patterned nanowire
two different pollution loads. Indoor. Air 9, electrodeposition. Nat. Mater. 5, 914–919.
165–179. 50. Ko, M., Aykanat, A., Smith, M.K., and Mirica,
K.A. (2017). Drawing sensors with ball-milled 63. Sun, Y.F., Liu, S.B., Meng, F.L., Liu, J.Y., Jin, Z.,
38. Yang, H., He, X.-W., Wang, F., Kang, Y., and blends of metal–organic frameworks and Kong, L.T., and Liu, J.H. (2012). Metal oxide
Zhang, J. (2012). Doping copper into ZIF-67 graphite. Sensors 17, 2192. nanostructures and their gas sensing
for enhancing gas uptake capacity and properties: a review. Sensors 12, 2610–2631.
visible-light-driven photocatalytic 51. Smith, M.K., and Mirica, K.A. (2017). Self-
degradation of organic dye. J. Mater. Chem. organized frameworks on textiles (SOFT): 64. Yao, M.S., Tang, W.X., Wang, G.E., Nath, B.,
22, 21849–21851. conductive fabrics for simultaneous sensing, and Xu, G. (2016). MOF thin film-coated metal
capture, and filtration of gases. J. Am. Chem. oxide nanowire array: significantly improved
39. Chen, E.X., Fu, H.R., Lin, R., Tan, Y.X., and Soc. 139, 16759–16767. chemiresistor sensor performance. Adv.
Zhang, J. (2014). Highly selective and sensitive Mater. 28, 5229–5234.
trimethylamine gas sensor based on cobalt 52. Pedersen, K.S., Perlepe, P., Aubrey, M.L.,
imidazolate framework material. ACS Appl. Woodruff, D.N., Reyes-Lillo, S.E., Reinholdt, 65. Tian, H., Fan, H., Li, M., and Ma, L. (2016).
Mater. Interfaces 6, 22871–22875. A., Voigt, L., Li, Z., Borup, K., Rouzières, M., Zeolitic imidazolate framework coated ZnO
et al. (2018). Formation of the layered nanorods as molecular sieving to improve
40. DMello, M.E., Sundaram, N.G., Singh, A., conductive magnet CrCl2(pyrazine)2 through selectivity of formaldehyde gas sensor. ACS
Singh, A.K., and Kalidindi, S.B. (2019). An redox-active coordination chemistry. Nat. Sens. 1, 243–250.
amine functionalized zirconium metal– Chem. 10, 1056–1061.
organic framework as an effective 66. DMello, M.E., Sundaram, N.G., and Kalidindi,
chemiresistive sensor for acidic gases. Chem. 53. Kung, C.W., Platero-Prats, A.E., Drout, R.J., S.B. (2018). Assembly of ZIF-67 metal–organic
Commun. 55, 349–352. Kang, J., Wang, T.C., Audu, C.O., Hersam, framework over tin oxide nanoparticles for
M.C., Chapman, K.W., Farha, O.K., and Hupp, synergistic chemiresistive CO2 gas sensing.
41. Long, J., Wang, S., Ding, Z., Wang, S., Zhou, J.T. (2018). Inorganic ‘‘conductive glass’’ Chemistry 24, 9220–9223.
Y., Huang, L., and Wang, X. (2012). Amine- approach to rendering mesoporous metal–
functionalized zirconium metal–organic 67. Drobek, M., Kim, J.H., Bechelany, M., Vallicari,
organic frameworks electronically conductive
framework as efficient visible-light C., Julbe, A., and Kim, S.S. (2016). MOF-based
and chemically responsive. ACS Appl. Mater.
photocatalyst for aerobic organic membrane encapsulated ZnO nanowires for
Interfaces 10, 30532–30540.
transformations. Chem. Commun. 48, 11656– enhanced gas sensor selectivity. ACS Appl.
11658. 54. Li, H., Wang, K., Sun, Y., Lollar, C.T., Li, J., and Mater. Interfaces 8, 8323–8328.
Zhou, H.-C. (2018). Recent advances in gas 68. Zhou, T., Sang, Y., Wang, X., Wu, C., Zeng, D.,
42. Sun, L., Campbell, M.G., and Dinc
a, M. (2016).
storage and separation using metal–organic and Xie, C. (2018). Pore size dependent gas-
Electrically conductive porous metal–organic
frameworks. Mater. Today 21, 108–121.
frameworks. Angew. Chem. Int. Ed. Engl. 55, sensing selectivity based on ZnO@ZIF
3566–3579. nanorod arrays. Sens. Actuators B Chem. 258,
55. Han, X., Godfrey, H.G.W., Briggs, L., Davies,
1099–1106.
43. Travlou, N.A., Singh, K., Rodrı́guez-Castellón, A.J., Cheng, Y., Daemen, L.L., Sheveleva,
E., and Bandosz, T.J. (2015). Cu–BTC MOF– A.M., Tuna, F., McInnes, E.J.L., Sun, J., et al. 69. Weber, M., Kim, J.H., Lee, J.H., Kim, J.Y.,
graphene-based hybrid materials as low (2018). Reversible adsorption of nitrogen Iatsunskyi, I., Coy, E., Drobek, M., Julbe, A.,
concentration ammonia sensors. J. Mater. dioxide within a robust porous metal–organic Bechelany, M., and Kim, S.S. (2018). High-
Chem. A 3, 11417–11429. framework. Nat. Mater. 17, 691–696. performance nanowire hydrogen sensors by
exploiting the synergistic effect of Pd
44. Feng, D., Lei, T., Lukatskaya, M.R., Park, J., 56. Liu, G., Chernikova, V., Liu, Y., Zhang, K.,
nanoparticles and metal–organic framework
Belmabkhout, Y., Shekhah, O., Zhang, C., Yi,
Huang, Z., Lee, M., Shaw, L., Chen, S., membranes. ACS Appl. Mater. Interfaces 10,
Yakovenko, A.A., Kulkarni, A., et al. (2018). S., Eddaoudi, M., and Koros, W.J. (2018).
34765–34773.
Robust and conductive two-dimensional Mixed matrix formulations with MOF
metal–organic frameworks with exceptionally molecular sieving for key energy-intensive 70. Rodenas, T., Luz, I., Prieto, G., Seoane, B.,
high volumetric and areal capacitance. Nat. separations. Nat. Mater. 17, 283–289. Miro, H., Corma, A., Kapteijn, F., Llabrés I
Energy 3, 30–36. Xamena, F.X.L.i., and Gascon, J. (2015).
57. Penner, R.M. (2017). A nose for hydrogen gas: Metal–organic framework nanosheets in
45. Sheberla, D., Bachman, J.C., Elias, J.S., Sun, fast, sensitive H2 sensors using polymer composite materials for gas
C.J., Shao-Horn, Y., and Dinc
a, M. (2017). electrodeposited nanomaterials. Acc. Chem. separation. Nat. Mater. 14, 48–55.
Conductive MOF electrodes for stable Res. 50, 1902–1910.
supercapacitors with high areal capacitance. 71. Xia, W., Mahmood, A., Zou, R., and Xu, Q.
Nat. Mat. 16, 220–224. 58. Yang, F., Kung, S.C., Cheng, M., Hemminger, (2015). Metal–organic frameworks and their
J.C., and Penner, R.M. (2010). Smaller is faster derived nanostructures for electrochemical
46. Talin, A.A., Centrone, A., Ford, A.C., Foster, and more sensitive: the effect of wire size on energy storage and conversion. Energy
M.E., Stavila, V., Haney, P., Kinney, R.A., the detection of hydrogen by single Environ. Sci. 8, 1837–1866.
Szalai, V., El Gabaly, F., Yoon, H.P., et al. palladium nanowires. ACS Nano 4, 5233–
(2014). Tunable electrical conductivity in 5244. 72. Li, X., Zheng, S., Jin, L., Li, Y., Geng, P., Xue,
metal-organic framework thin-film devices. H., Pang, H., and Xu, Q. (2018). Metal–organic
Science 343, 66–69. 59. Favier, F., Walter, E.C., Zach, M.P., Benter, T., framework-derived carbons for battery
and Penner, R.M. (2001). Hydrogen sensors applications. Adv. Energy Mater. 8, 1800716.
47. Campbell, M.G., Liu, S.F., Swager, T.M., and and switches from electrodeposited
Dinc
a, M. (2015). Chemiresistive sensor arrays palladium mesowire arrays. Science 293, 73. Kim, C.R., Uemura, T., and Kitagawa, S. (2016).
from conductive 2D metal–organic 2227–2231. Inorganic nanoparticles in porous
frameworks. J. Am. Chem. Soc. 137, 13780– coordination polymers. Chem. Soc. Rev. 45,
13783. 60. Nyberg, C., and Tengstål, C.G. (1984). 3828–3845.
Adsorption and reaction of water, oxygen,
48. Smith, M.K., Jensen, K.E., Pivak, P.A., and and hydrogen on Pd (100): identification of 74. Koo, W.-T., Cha, J.-H., Jung, J.-W., Choi, S.-J.,
Mirica, K.A. (2016). Direct self-assembly of adsorbed hydroxyl and implications for the Jang, J.-S., Kim, D.-H., and Kim, I.-D. (2018).
conductive nanorods of metal–organic catalytic H2–O2 reaction. J. Chem. Phys. 80, Few-layered WS2 nanoplates confined in Co,
frameworks into chemiresistive devices on 3463–3468. N-doped hollow carbon nanocages: abundant

Chem 5, 1–26, August 8, 2019 25


Please cite this article in press as: Koo et al., Metal-Organic Frameworks for Chemiresistive Sensors, Chem (2019), https://doi.org/10.1016/
j.chempr.2019.04.013

WS2 edges for highly sensitive gas sensors. 86. Koo, W.T., Yu, S., Choi, S.J., Jang, J.S., mesoporous In/Ga oxides and their ultra-
Adv. Funct. Mater. 28, 1802575. Cheong, J.Y., and Kim, I.D. (2017). Nanoscale sensitive ethanol-sensing properties.
PdO catalyst functionalized Co3O4 hollow J. Mater. Chem. A 6, 14930–14938.
75. Ko, K.Y., Song, J.G., Kim, Y., Choi, T., Shin, S., nanocages using MOF templates for selective
Lee, C.W., Lee, K., Koo, J., Lee, H., Kim, J., detection of acetone molecules in exhaled 96. Li, W., Wu, X., Liu, H., Chen, J., Tang, W., and
et al. (2016). Improvement of gas-sensing breath. ACS Appl. Mater. Interfaces 9, 8201– Chen, Y. (2015). Hierarchical hollow ZnO
performance of large-area tungsten disulfide 8210. cubes constructed using self-sacrificial ZIF-8
nanosheets by surface functionalization. ACS frameworks and their enhanced benzene gas-
Nano 10, 9287–9296. 87. Yang, D.J., Kamienchick, I., Youn, D.Y., sensing properties. New J. Chem. 39, 7060–
Rothschild, A., and Kim, I.-D. (2010). 7065.
76. Indra, A., Song, T., and Paik, U. (2018). Metal Ultrasensitive and highly selective gas sensors
organic framework derived materials: based on electrospun SnO2 nanofibers
progress and prospects for the energy 97. Jo, Y.M., Kim, T.H., Lee, C.S., Lim, K., Na,
modified by Pd loading. Adv. Funct. Mater. C.W., Abdel-Hady, F., Wazzan, A.A., and Lee,
conversion and storage. Adv. Mater. 30, 20, 4258–4264.
1705146. J.H. (2018). Metal–organic framework-derived
88. Koo, W.T., Choi, S.J., Jang, J.S., and Kim, I.D. hollow hierarchical Co3O4 nanocages with
77. Wang, X.-F., Song, X.-Z., Sun, K.-M., Cheng, (2017). Metal-organic framework templated tunable size and morphology: ultrasensitive
L., and Ma, W. (2018). MOFs-derived porous synthesis of ultrasmall catalyst loaded ZnO/ and highly selective detection of
nanomaterials for gas sensing. Polyhedron ZnCo2O4 hollow spheres for enhanced gas methylbenzenes. ACS Appl. Mater. Interfaces
152, 155–163. sensing properties. Sci. Rep. 7, 45074. 10, 8860–8868.

78. Lü, Y., Zhan, W., He, Y., Wang, Y., Kong, X., 89. Jang, J.S., Koo, W.T., Choi, S.J., and Kim, I.D. 98. Koo, W.T., Jang, J.S., Choi, S.J., Cho, H.J.,
Kuang, Q., Xie, Z., and Zheng, L. (2014). MOF- (2017). Metal organic framework-templated and Kim, I.D. (2017). Metal–organic
templated synthesis of porous Co3O4 chemiresistor: sensing type transition from framework templated catalysts: dual
concave nanocubes with high specific surface P-to-N Using hollow metal oxide polyhedron sensitization of PdO–ZnO composite on
area and their gas sensing properties. ACS via galvanic replacement. J. Am. Chem. Soc. hollow SnO2 nanotubes for selective acetone
Appl. Mater. Interfaces 6, 4186–4195. 139, 11868–11876. sensors. ACS Appl. Mater. Interfaces 9,
18069–18077.
79. Wang, J.-L., Zhai, Q.-G., Li, S.-N., Jiang, Y.-C.,
90. Oh, M.H., Yu, T., Yu, S.H., Lim, B., Ko, K.T.,
and Hu, M.-C. (2016). Mesoporous In2O3
Willinger, M.G., Seo, D.H., Kim, B.H., Cho, 99. Zhang, H., Liu, G., Shi, L., and Ye, J. (2018).
materials prepared by solid-state thermolysis
M.G., Park, J.H., et al. (2013). Galvanic Single-atom catalysts: emerging
of indium-organic frameworks and their high
replacement reactions in metal oxide multifunctional materials in heterogeneous
HCHO-sensing performance. Inorg. Chem.
nanocrystals. Science 340, 964–968. catalysis. Adv. Energy Mater. 8.
Commun. 63, 48–52.
91. Koo, W.T., Choi, S.J., Kim, S.J., Jang, J.S.,
80. Gao, P., Liu, R., Huang, H., Jia, X., and Pan, H. 100. Park, J.G., Aubrey, M.L., Oktawiec, J.,
Tuller, H.L., and Kim, I.D. (2016).
(2016). MOF-templated controllable synthesis Chakarawet, K., Darago, L.E., Grandjean, F.,
Heterogeneous sensitization of metal–
of a-Fe2O3 porous nanorods and their gas Long, G.J., and Long, J.R. (2018). Charge
organic framework driven metal@metal oxide
sensing properties. RSC Adv. 6, 94699–94705. delocalization and bulk electronic
complex catalysts on an oxide nanofiber
conductivity in the mixed-valence metal–
81. Li, W., Wu, X., Han, N., Chen, J., Qian, X., scaffold toward superior gas sensors. J. Am.
organic framework Fe(1,2,3-triazolate)2(BF4)x.
Deng, Y., Tang, W., and Chen, Y. (2016). MOF- Chem. Soc. 138, 13431–13437.
J. Am. Chem. Soc. 140, 8526–8534.
derived hierarchical hollow ZnO nanocages
with enhanced low-concentration VOCs gas- 92. Rui, K., Wang, X., Du, M., Zhang, Y., Wang, Q.,
Ma, Z., Zhang, Q., Li, D., Huang, X., Sun, G., 101. Stassen, I., Styles, M., Grenci, G., Gorp, H.V.,
sensing performance. Sens. Actuators B Vanderlinden, W., Feyter, S.D., Falcaro, P.,
Chem. 225, 158–166. et al. (2018). Dual-function metal–organic
framework-based wearable fibers for gas Vos, D.D., Vereecken, P., and Ameloot, R.
82. Jang, J.-S., Koo, W.-T., Kim, D.-H., and Kim, probing and energy storage. ACS Appl. (2016). Chemical vapour deposition of zeolitic
I.-D. (2018). In situ coupling of Mater. Interfaces 10, 2837–2842. imidazolate framework thin films. Nat. Mater.
multidimensional MOFs for heterogeneous 15, 304–310.
metal-oxide architectures: toward sensitive 93. Choi, S.J., Choi, H.J., Koo, W.T., Huh, D., Lee,
chemiresistors. ACS Cent. Sci. 4, 929–937. H., and Kim, I.D. (2017). Metal–organic 102. Ma, X., Kumar, P., Mittal, N., Khlyustova, A.,
framework-templated PdO-Co3O4 Daoutidis, P., Mkhoyan, K.A., and Tsapatsis,
83. Miller, D.R., Akbar, S.A., and Morris, P.A. nanocubes functionalized by SWCNTs: M. (2018). Zeolitic imidazolate framework
(2014). Nanoscale metal oxide-based improved NO2 reaction kinetics on flexible membranes made by ligand-induced
heterojunctions for gas sensing: a review. heating film. ACS Appl. Mater. Interfaces 9, permselectivation. Science 361, 1008–1011.
Sens. Actuators B Chem. 204, 250–272. 40593–40603.
103. Wang, A., Li, J., and Zhang, T. (2018).
84. Yu, J., Mu, C., Yan, B., Qin, X., Shen, C., Xue, 94. Wang, Y., Lü, Y., Zhan, W., Xie, Z., Kuang, Q., Heterogeneous single-atom catalysis. Nat.
H., and Pang, H. (2017). Nanoparticle/MOF and Zheng, L. (2015). Synthesis of porous Rev. Chem. 2, 65–81.
composites: preparations and applications. Cu2O/CuO cages using Cu-based metal–
Mater. Horiz. 4, 557–569. organic frameworks as templates and their
104. Kim, S.-J., Choi, S.-J., Jang, J.-S., Cho, H.-J.,
gas-sensing properties. J. Mater. Chem. A 3,
85. Chen, L., Luque, R., and Li, Y. (2017). Koo, W.-T., Tuller, H.L., and Kim, I.-D. (2017).
12796–12803.
Controllable design of tunable exceptional high-performance of Pt-based
nanostructures inside metal–organic 95. Cui, Y.F., Jiang, W., Liang, S., Zhu, L.F., and bimetallic catalysts for exclusive detection of
frameworks. Chem. Soc. Rev. 46, 4614–4630. Yao, Y.J. (2018). MOF-derived synthesis of exhaled biomarkers. Adv. Mater. 29, 1700737.

26 Chem 5, 1–26, August 8, 2019

Вам также может понравиться