You are on page 1of 9

www.acsami.

org Research Article

Ultrafast Degradation and High Adsorption Capability of a Sulfur


Mustard Simulant under Ambient Conditions Using Granular UiO-
66-NH2 Metal−Organic Gels
Chuan Zhou, Bo Yuan, Shouxin Zhang, Guang Yang, Lin Lu,* Heguo Li,* and Cheng-an Tao*
Cite This: ACS Appl. Mater. Interfaces 2022, 14, 23383−23391 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: Zirconium-based metal−organic frameworks (Zr-MOFs) have been


considered as prospective materials for the degradation of nerve chemical warfare agents
(CWAs) but show poor catalytic performance toward blister agents. Moreover, the
powder issues and the poor adsorption capability also remain as the major challenges for
Downloaded via KOREA UNIV on April 8, 2023 at 20:26:44 (UTC).

the application of Zr-MOFs in practical CWA detoxification. Herein, a series of defected


granular UiO-66-NH2 metal−organic gels are synthesized via adjusting the amount of
added concentrated hydrochloric acid for the decontamination of 2-chloroethyl ethyl
sulfide (2-CEES), a sulfur mustard simulant. The half-life of 2-CEES decontaminated by
defected granular UiO-66-NH2 metal−organic gels can be shortened to 7.6 min, which
is the highest reported value for MOFs under ambient conditions. The mechanism of
decontamination is that the amino group on the linkers in UiO-66-NH2 MOGs
undergoes a substitution reaction with 2-CEES to yield 2-(2-(ethylthio)ethylamino)-
terephthalic acid, which is less toxic and fixed in the frameworks. The recycling test corroborates that the granular UiO-66-NH2
xerogels possess good stability and reusability. Static adsorption and desorption tests show that UiO-66-NH2 xerogels possess a high
2-CEES vapor adsorption capacity of 802 mg/g after exposure for 1 d and only 28 wt % desorption capacity after air exposure for 7
d. The dual function of ultrafast degradation and high adsorption capability provide a firm foundation for using UiO-66-NH2
xerogels as a future protection media.
KEYWORDS: metal−organic gel, metal−organic framework, chemical warfare agent, 2-chloroethyl ethyl sulfide, detoxification, adsorption

1. INTRODUCTION MOFs are unsuitable for decontamination application because


Chemical warfare agents (CWAs) including nerve agents and of their low moisture stability and high cost.13
blistering agents continue to be a significant long-lasting threat Zirconium-based metal−organic frameworks (Zr-MOFs) are
to both mankind and environment, although prohibited by an foreseen as promising materials for the degradation of CWAs
international convention. Sulfur mustard (HD) is a canonical due to their ease and plentiful synthesis methods and
vesicant that can cause severe erosion of epithelial tissue, injury exceptional thermal, chemical, and mechanical stability.14−16
of the respiratory/digestive tract, and damage of visual systems UiO-66-NH2 is a representative type of Zr-MOF that is
on contact.1−3 Due to the facile synthesis, high stability, and assembled by Zr6(μ3-O)4(μ3-OH)4 clusters with 2-amino-
low volatility, the production, storage, and distribution of HD terephthalic acid through coordination bonds. It exhibits
still exist to date.4,5 As a result, the development of advanced excellent activity on actual nerve agents and their simulants but
materials for rapid detoxification of HD into less toxic products lower efficacy against the detoxification of blistering agents and
remains to be a momentous concern. their simulants.17−20 Furthermore, the crystallization of UiO-
Recently, metal−organic frameworks (MOFs) have gained 66-NH2 tends to form microcrystalline powders. The
tremendous research attention as promising candidates for practicability of powdered MOFs is greatly restricted due to
CWA adsorption and decontamination owing to their its particle aggregation, poor processability, mass-transfer
adjustable structures and functionalities, well-defined porous limitations, high air resistance, and difficult recycle.21−24 On
structures, large surface area, and rich catalytic sites.6−8 These
isolated metal sites within nanoporosity endow MOFs with Received: February 8, 2022
strong reactivity toward CWAs. Although multifarious MOFs Accepted: May 2, 2022
have been proved to effectively detoxify HD in aqueous Published: May 12, 2022
solution,9,10 the poor aqueous solubility of HD and its normal
presence in the gaseous phase improve the difficulty of HD
decontamination in real-life scenarios.11,12 Moreover, most

© 2022 American Chemical Society https://doi.org/10.1021/acsami.2c02401


23383 ACS Appl. Mater. Interfaces 2022, 14, 23383−23391
ACS Applied Materials & Interfaces www.acsami.org Research Article

Scheme 1. Schematic Illustration of the Formation of UiO-66-NH2 Xerogel and the Decontamination and Adsorption Process
of 2-CEES

the other hand, adsorption performance is another important use of MOGs to ultrafast decontaminate sulfur mustard
capability for defense against CWAs. However, due to the simulant in the pure phase and the first example of MOGs with
small size of the internal pore size of the most MOFs, it is dual function of decontamination and adsorption.
difficult for many large-sized CWA molecules to enter MOFs’
interior, resulting in poor adsorption performance. 2. EXPERIMENTAL SECTION
Metal−organic gels (MOGs), known as a class of extended
2.1. Materials. Zirconyl chloride octahydrate (ZrOCl2·8H2O,
MOF architecture, have aroused extensive concern because of 99.9%), hydrofluoric acid (HF, 49%), dimethyl sulfoxide-d6 (DMSO-
their low density, large surface area, open metal sites, abundant d6, 99.9%), and deuterium oxide (D2O, 99.9%) are obtained from
pore structure, gentle synthesis condition, and low cost of Macklin Biochemical (Shanghai, China). 2-Aminoterephthalicacid
production.25−29 MOGs can be manufactured as macroscopic (NH2-BDC, 99%) is purchased from Alfa Aesar Chemical (Tianjin,
monoliths or granules by aggregating MOF particles into China). Anhydrous ethanol (99.8%) and N,N-dimethylformamide
tridimensional networks through metal−ligand coordination (DMF, 99.9%) are obtained from Aladdin Chemical (Shanghai,
associated with van der Waals interactions, hydrogen binding, China). Potassium phosphate (K3PO4, 99%) and methanol (99.8%)
and π−π stacking.30,31 The critical point is to adjust synthetic are purchased from J&K Chemical (Chongqing, China). Acetone and
concentrated hydrochloric acid (HCl, 37%) are obtained from
factors such as reactant concentration, temperature, time, and Sinopharm Chemical (China). 2-CEES (98%) and 2-hydroxydiethyl
solvent for the preparation of MOGs, which consisted of sulfide (2-HEES, 98%) are purchased from Inno Chemical (Beijing,
nanoparticles.32−34 Compared to MOFs, MOGs can not only China). All chemicals are used as received without further
surmount the limitations of MOF powders in practical purification.
applications but also their hierarchically porous structure is 2.2. Preparation of Defected Granular Zr-MOGs. The
conducive to reducing diffusion barriers between substrates defected granular Zr-MOGs are acquired with different amounts of
and active sites, and accelerating the mass-transfer rate, which concentrated hydrochloric acid (HCl), which are labeled Zr-MOG-X
endow MOGs with reinforced performance in adsorption and (where x = 0, 1, 4, 7, and 10), representing the volume of HCl added
to the reaction. Typically, ZrOCl2·8H2O (2.34 g, 7.25 mmol) and
catalysis; moreover, the hierarchically porous structure also NH2-BDC (0.92 g, 5 mmol) are weighed into a 100 ml Pyrex Schott
endows MOGs with special adsorption capacity, which makes bottle and dissolved in DMF (30 mL) under sonication. Then, a
MOGs show good adsorption performance for CWAs. Very certain amount of HCl is added into the solution. Subsequently, the
recently, the macroscopic monolithic UiO-66-X xerogels for bottle is placed in an oven after ultrasonic dispersion for 5 min. The
decomposition of CWAs in solution have been reported by us reaction temperature is held at 100 °C for 2 h. After cooling naturally,
for the first time.35 However, to the best of our knowledge, the synthesized gels are washed three times with anhydrous ethanol
there have been investigations neither about the decontami- and dried in air at 30 °C. The obtained products are further washed
nation of HD on MOGs in the pure phase nor the dual three times in acetone and methanol under sonication and then dried
at 30 °C overnight to produce a series of defected Zr-MOG-X, as
function of decontamination and adsorption toward HD or its shown in Scheme 1. Zr-MOG-X is activated by heating to 110 °C
simulants. under vacuum for 12 h prior to use.
Herein, a series of defected granular UiO-66-NH2 MOGs 2.3. Degradation of 2-CEES. The “pure-phase” detoxification of
are prepared by using HCl as both modulator and activator 2-CEES in this paper indicates the direct destruction of 2-CEES by
(Scheme 1). The effect of the modulator on structure, Zr-MOG-X without any solvents or additives. 2-CEES (1 μL) is
degradation ability, and adsorption capability toward 2- spiked into 10 mg of desired Zr-MOG-X in a 4 mL vial and allows to
chloroethyl ethyl sulfide (2-CEES), a sulfur mustard simulant, stand for various periods of time before extracting the residual 2-
in the pure phase (without any solvents or additives) under CEES from the vial with acetonitrile. Specifically, 3 mL of acetonitrile
is added into the above vial and vigorously mixed with a shaker for 3
ambient conditions is investigated. The synthesized UiO-66- min. Subsequently, the extracted solution is passed through a 0.22 μm
NH2 xerogels are demonstrated to not only exhibit high filter using plastic syringes prior to GC analysis.
decontamination performance for 2-CEES and good reusability 2.4. Adsorption and Desorption of 2-CEES. Closed vapor
but also show high adsorption capability toward 2-CEES. To adsorption. Three pieces of glass sheet with a diameter of 2 cm
the best of our knowledge, this is the first demonstration of the containing about 30 mg of Zr-MOG-X were placed into a 200 mL

23384 https://doi.org/10.1021/acsami.2c02401
ACS Appl. Mater. Interfaces 2022, 14, 23383−23391
ACS Applied Materials & Interfaces www.acsami.org Research Article

Figure 1. SEM images and particle size distributions of (a) Zr-MOG-0, (b) Zr-MOG-1, (c) Zr-MOG-4, (d) Zr-MOG-7, and (e) Zr-MOG-10.
Insets: Photographs of the corresponding Zr-MOG-X.

Figure 2. (a) PXRD patterns, (b) TGA curves, (c) FT-IR spectra, and (d) N2 adsorption−desorption isotherms of the as-synthesized Zr-MOG-X.

reaction vessel. 2-CEES (250 μl) was placed in a 1 mL vial located USA) under air flow at a heating rate of 5 °C/min up to 800 °C.
inside the reaction vessel. Then, the vessel was immediately sealed Nitrogen adsorption−desorption isotherms were measured at 77 K on
using a ground glass stopper and kept at 25 °C for 1 day. Finally, the
a Micrometer ASAP 2020 analyzer (USA). Nuclear magnetic
vessel was opened, and the xerogels were weighed. The mass obtained
as a result of the interactions of 2-CEES vapors with the materials was resonance (NMR) spectra were obtained from a Bruker 400M
recorded and expressed as weight uptake (WU, mg/g). spectrometer (Germany). Fourier transform infrared (FT-IR) spectra
Desorption. After the vessel was opened, the materials were left were recorded on a spectrometer (PerkinElmer Frontier, USA). Gas
inside a fume cupboard under ambient conditions, and after 2 and 7
days, they were weighed. A desorbed amount was calculated as a mass chromatography (GC) with a flame photometric detector (Agilent
difference of the materials before and after the desorption process. Technologies) was used for quantitative analysis of residual 2-CEES
2.5. Characterization. A Zeiss Sigma 300 (Germany) field- after the decomposition experiments (Agilent Technologies 6890,
emission scanning electron microscope was used for morphological
USA). Gas chromatography−mass spectrometry (GC−MS, Agilent
studies. PXRD patterns were collected on a diffractometer (Rigaku D/
MAX 2500V, Japan) with Cu Kα = 1.5406 Å. Thermogravimetric Technologies 7890A, USA) with a 5975C detector (Agilent
analyses (TGA) were performed on a thermal analyzer (Q600 SDT, Technologies) was used to analyze the degradation products.

23385 https://doi.org/10.1021/acsami.2c02401
ACS Appl. Mater. Interfaces 2022, 14, 23383−23391
ACS Applied Materials & Interfaces www.acsami.org Research Article

Figure 3. (a) Profiles of the removal efficiency. (b) Half-life of Zr-MOG-X. (c) Profiles of the removal efficiency after degradation of Zr-MOG-4
toward 2-CEES. (d) Proposed decontamination reaction of Zr-MOG-X and 2-CEES.

3. RESULTS AND DISCUSSION with the increase of the modulator, indicating that larger
3.1. Characterization of UiO-66-NH2 MOGs. Concen- crystallites are obtained. These results are also consistent with
trated hydrochloric acid (HCl, 37%) possesses functions of the SEM data.
both modulation and activation.36,37 We adopt HCl as both a TGA analyses are performed to investigate the information
modulator and an activator to synthesize a series of defected of missing linker defects of the as-synthesized Zr-MOG-X
granular Zr-MOG-X xerogels by adjusting the amount of HCl. samples, as shown in Figure 2b. The missing linker defects of
As shown in the insets of Figure 1, as HCl increases, the color Zr-MOG-X can be calculated by Lillerud’s protocol.38 The
of macroscopic monolithic Zr-MOG-X (∼ a few millimeters) estimated numbers of the resulting missing linker of Zr-MOG-
changes from brown to light yellow. The surface morphology 0, Zr-MOG-1, Zr-MOG-4, Zr-MOG-7, and Zr-MOG-10 are
and particle size of the as-synthesized Zr-MOG-X xerogels are 1.04, 1.87, 1.70, 1.50, and 1.00, respectively. The molar ratios
investigated by SEM observation (Figure 1). The SEM images of HCOO− and NH2-BDC are obtained from 1H NMR spectra
show the rough surface of all Zr-MOG-X samples, which are (Table S1 and Figure S1-S7). Based on above both TGA and
1
primarily attributed to the densely packed nanoparticles. As H NMR data, the composition of Zr-MOG-X can be
shown in Figure 1a, the Zr-MOG-0 xerogel synthesized calculated, and the results are listed in Table S2. On the
without a modulator has smallest particle size and exhibit an basis of crystal growth theory, Zr-MOG-0 is expected to
aggregated and irregularly interconnected structure composed possess smaller particle size and higher defect density on
of nanocrystals. As hydrochloric acid continues to increase, we account of the faster rate of crystal nucleation. Contrary to
can clearly see that the borders of the nanoparticles that made previous reports,39,40 we astonishingly find that with the
up the samples become clearer and the sizes of the further increase of the modulator, fewer defects are generated
nanoparticles grow larger. Obviously, the particle sizes of the due to the enhancement of the crystallinity for the Zr-MOG-X
Zr-MOG-X xerogels can be adjusted by the addition of a (X = 1, 4, 7, and 10), which correspond well with the PXRD
modulator. data.
The crystalline characteristic of the synthesized Zr-MOG-X The FT-IR spectrum is further performed to substantiate the
products is demonstrated by PXRD measurements (Figure 2a). consequence of the prepared xerogels (Figure 2c). The FTIR
It can be observed that there are obvious peaks at 7.3 and 8.5° spectra of all the samples are consistent with one another. The
of the prepared Zr-MOG-X, broadly consistent with the overall resultant Zr-MOG-X delivers the −NH2 asymmetric
simulated theoretical PXRD patterns. The PXRD pattern of and symmetric stretching characteristic peaks at 3468 and
Zr-MOG-0 shows the broadness and weakness of the 3356 cm−1, the two distinctive peaks at 1567 and 1384 cm−1 of
characteristic diffraction peaks, which indicate that Zr-MOG- carboxyl groups, and the Zr−O stretching vibration peaks in
0 is of low crystallinity with the amorphous phase during the the range of at 800−600 cm−1, in good agreement with
gelation process of UiO-66-NH2 primary nanoparticles. The previous reports.41 By comparing these samples, it can be
diffraction peaks of the fabricated samples become sharper found that the intensity of peaks at 1258 cm−1 corresponding
23386 https://doi.org/10.1021/acsami.2c02401
ACS Appl. Mater. Interfaces 2022, 14, 23383−23391
ACS Applied Materials & Interfaces www.acsami.org Research Article

Table 1. Comparison of Various Materials Used as Protection Media for the Detoxification of HD or Its Simulant 2-CEES
materials material weight, mg material amount, μmol agent volume, μL half-life refs
MgO 111 2754.3 HD, 5 17.8 h 50
CaO 115 2050.6 HD, 5 8.7 h 51
Al2O3 80 784.6 HD, 4 6.3 h 52
TiO2 200 2504.1 HD, 5 3.8 h 53
MnO2 50 575.1 2-CEES, 2 4.6 h 54
VOx 200 2007.2 HD, 5 6.3 h 55
NiO NpsAg-clinoptilolite zeolite 300 2-CEES, 8 7.3 h 56
POMs/Carbon 100 HD, 1 15 h 57
POMs/Al2O3 100 HD, 5 214 min 58
HKUST-1 50 75.9 2-CEES, 5 16.1 min 59
Ag NPs@HKUST-1 30 HD, 4 15 min 60
UiO-66 40 24.0 2-CEES, 6 1.3 h 61
UiO-66-NH2@PAN-2h 20 2-CEES, 5 <0.5 h 20
UiO-66-NH2@PAN-3h 20 2-CEES, 5 <0.5 h 20
Zr-MOG-0 10 5.7 2-CEES, 1 8.2 min this work
Zr-MOG-1 10 5.7 2-CEES, 1 7.6 min this work
Zr-MOG-4 10 5.7 2-CEES, 1 12.6 min this work
Zr-MOG-7 10 5.7 2-CEES, 1 25.4 min this work
Zr-MOG-10 10 5.7 2-CEES, 1 33.7 min this work

to the δ(OH)+ δ(CH) mode weakens gradually with the nearly 80% conversion within 30 min, which may be attributed
increase of the modulator as fewer defects are introduced. to the smaller particle size of the samples.
The N2 adsorption−desorption isotherms of Zr-MOG-X are The stability and reusability are important properties of
displayed in Figure 2d. Each of the samples manifests a decontaminants. The stability of Zr-MOG-X (X = 4, 7, or 10)
representative IV hysteresis loop, which reveals the existence of after the decontamination of 2-CEES is valuated by PXRD and
micro-mesoporosity in the structure of gels. As shown in Table N2 adsorption−desorption isotherms (Figure S10). Signifi-
S3, with the increasing amount of modulator, the Brunauer− cantly and impressively, the PXRD patterns of the collected Zr-
Emmett−Teller (BET) surface area of Zr-MOG-X is first MOG-X show that the crystal structure of Zr-MOG-X can be
significantly increased to 1178 m2 g−1 and then slightly essentially maintained. Furthermore, the N2 adsorption−
decreased to 1117 m2 g−1. The same tendency as the BET desorption isotherms indicate that Zr-MOG-X exhibits the
surface area is observed with the pore volume. Furthermore, same type IV behavior as the pristine Zr-MOG-X and show an
the NLDFT and BJH pore diameter distributions are displayed almost constant surface area. Therefore, Zr-MOG-X can be
in Figures S8 and S9, respectively. It can be obviously observed considered as a renewable decontaminant suitable for 2-CEES
that both the strength of peaks for NLDFT pore diameter degradation. A reusability test is further carried out to
distributions and the pore diameter for BJH pore diameter demonstrate. The recyclable Zr-MOG-4 displays a very similar
distributions are increased. Therefore, based on the above conversion profiles to fresh Zr-MOG-4 in the beginning and
results, it can be reconfirmed that the modulator (HCl) further a little lower than that of fresh Zr-MOG-4, as presented
influences the growth of the MOF during the progress of self- in Figure 3c, suggesting the preservation of most decontami-
nation performance. Both these excellent stability and
assembly of the infrastructure architecture building blocks.
reusability further ensure defected granular UiO-66-NH2
3.2. Detoxification Performance toward 2-CEES. The
metal−organic gels for CWA removal as a promising
decontamination performance of the defected granular Zr-
candidate.
MOG-X is investigated by conducting the detoxification of the
In order to investigate the mechanism of degradation of 2-
sulfur mustard simulant 2-CEES in the pure phase at room CEES by Zr-MOG-X, it is necessary to validate the
temperature. The conversion profiles of the as-synthesized Zr- decomposition products extracted with acetonitrile by GC−
MOG-X MOGs are presented in Figure 3a as a function of MS analysis. Take Zr-MOG-4 as an example (Figure S11), the
time. The results demonstrate that initial degradation reactions data illustrates that 2-hydroxyethyl ethyl sulfide (2-HEES)
are faster, followed by a steady-state reaction. The fast initial (with m/z values at 106, 75, and 47) as the sole hydrolysis
reaction is due to the rapid adsorption and diffusion of 2-CEES product of 2-CEES, is monitored. This hydrolysis product with
within the pores and its interaction with the accessible reactive relatively low toxicity is usually the goal of chemical
sites. The restricted surface reaction prevails over the initial fast decontamination of 2-CEES. Similar results are also observed
reaction when the reactive sites are depleted. The Zr-MOG-0 by other researchers in several studies.20,42,43 Unfortunately,
and Zr-MOG-1 samples show outstanding decontamination the amount of the hydrolysis product is extremely low. In
performance toward 2-CEES with half-lives of about 8 min. addition to the hydrolysis product, the residual 2-CEES
The decontamination performance of Zr-MOG-X gradually amount (with m/z values at 124, 109, 89, 75, and 47) is more.
deteriorates as the increase of the amount of HCl added during According to the GC analysis, about 83% of 2-CEES has been
synthesis. The values of half-life for all the Zr-MOG-X degraded. To find out the degradation products, the residual
materials are shown in Figure 3b. Specially, Zr-MOG-0 with Zr-MOG-4 solid is digested with 750 μL of 1.8 M HF in
unsatisfactory porosity and missing linker defects exhibits DMSO-d6. As shown in Figure S12, the unperturbed 2-amino
fastest 2-CEES degradation rate with a half-life of 8.2 min and terephthalic acid linker and the 2-(2-(ethylthio)ethylamino)-
23387 https://doi.org/10.1021/acsami.2c02401
ACS Appl. Mater. Interfaces 2022, 14, 23383−23391
ACS Applied Materials & Interfaces www.acsami.org Research Article

Figure 4. 2-CEES adsorption capacity of Zr-MOG-X at 298 K (a) and the amounts desorbed expressed as weight change of the samples exposed to
air for 2 d and 7 d (b).

terephthalic acid reaction product were observed. The reaction literature is usually about 100 mg (∼2000 μmol vs 4−5 μL
is shown in Figure 3d. The ratio of these two species is agent) and the values of the half-life are generally in the range
determined by the ratio of H atoms at 7.79 and 7.86 ppm. As of the tens of minutes to several hours. Notably, Zr-MOG-X
listed in Table S4, the reacted linkers achieve 19% among all can be used directly for 2-CEES decontamination with less
linkers, that is, 1.14 per 6 linkers. Considering that the Zr- amount (10 mg vs 1 μL 2-CEES) and lower half-life (about 8
MOG-4 MOG has the linker missing defects, there is nearly min), highlighting the advanced degradation of our developed
one reacted 2-CEES molecule in each periodic pores. Zr-MOG-X materials.
Assuming that the Zr-MOG-4 MOG has the ideal formula of 3.3. Adsorption Capability of Zr-MOG-X Xerogels.
UiO-66-NH2, the amount of reacted 2-CEES can be estimated The adsorption capabilities of all Zr-MOG-X xerogels against
to be 6.5 μmol, which is about 76 mol % of original 2-CEES. It the vapors of 2-CEES are evaluated based on the WUs after the
can be concluded that most of the 2-CEES is decontaminated exposure. The results are presented in Figure 4a. The
by reaction with amino of linkers, and a very small part is adsorption capacity of Zr-MOG-X has been significantly
hydrolyzed to 2-HEES. improved with the increasing amount of the modulator, but
Previous reports44−49 have revealed that the two factors of a decrease is observed by further increasing the amount of
the size of MOFs and defect degree are often intertwined to modulator after a 1 day exposure. This trend indicates the
affect the decontamination performance of MOF powders in trade-offs between adsorption capacity and material formula-
solution. The smaller the size of MOF crystals is, the more tions. Interestingly, Zr-MOG-0 shows outstanding degradation
specific surface area/active sites are available. In addition, the activity toward 2-CEES benefiting from its small particle size
greater the degree of MOF defects (such as missing linkers or but lower adsorption capacity due to its smaller pore volume.
nodes) is, the better the decontamination performance of Zr-MOG-4 reveals the highest WU value of 802 mg/g after 1
MOF is. The three-dimensional relationship diagram of the day of exposure, which demonstrates a much higher value than
half-life with the sizes of the MOF Zr-MOG-X MOGs and that of various active inorganic nanomaterials62−64 with WU
their defect degrees are shown in Figure S13. For the other values in the range of 90−210 mg/g and porous carbon textiles
MOGs (i.e., Zr-MOG-1, Zr-MOG-4, Zr-MOG-7, and Zr- under the same condition with a WU value of 255 mg/g.65 To
MOG-10), with the increase of the amount of HCl, the size of investigate whether MOG would degrade 2-CEES after
the obtained MOGs increased, the defect density also physical adsorption of 2-CEES, they were extracted and
increased, and the half-life of the degradation reaction quantified with GC. As shown in Figure S17, there were
decreased, which is consistent with the previous reports. neither an obvious peak of the hydrolyzed product, 2-HEES,
Besides, in the MOG case, the intergranular porosity is another and nor peaks of other substances, and 2-CEES was quantified
key factor for the catalysis performance. We further investigate to be 23.4 mg in total, which is almost equal to the value
the properties and the 2-CEES degradation performance of evaluated by weight. This indicates that the MOG hardly reacts
selected Zr-MOG-0 and Zr-MOG-4 xerogels and their with 2-CEES after adsorption of 2-CEES vapor, which is not
corresponding grinding powder. Compared with Zr-MOG-0 consistent with the situation of UiO-66-NH2 powders66 or
and Zr-MOG-4 xerogels, the porosity, crystallinity, and fabrics.20
degradation performance of milled Zr-MOG-0 and Zr-MOG- Another important feature of protection materials is their
4 powder decreased obviously (Figures S14−S16). This can be ability to retain adsorbed CWA on their surface/in the porous
explained by the fact that mechanical grinding destroys part of structure. Though there have been few reports on the
the porous structure of xerogel, resulting in decreased porosity adsorption of CWA vapors by MOFs,42,67 the evaluation of
and crystallinity and increased mass-transfer resistance. desorption process related to two critical factors for real
In order to evaluate the degradation performance of the as- applications including the released absolute amount and the
synthesized Zr-MOG-X MOGs, a comparison is made between percent of the desorbed amount of toxic vapors has not been
their performance and that of some typical reference materials involved yet. The weight change of the Zr-MOG-X exposed to
(metal oxides, zeolites, POMs, and MOFs), as shown in Table 2-CEES vapors for 1 day is monitored after the air desorption
1. It should be noted that the amount of these reference tests for 2 d and 7 d, respectively. The absolute values of
materials used for HD or 2-CEES decontamination in the weight losses are included in Figure 4a, and the weights of Zr-
23388 https://doi.org/10.1021/acsami.2c02401
ACS Appl. Mater. Interfaces 2022, 14, 23383−23391
ACS Applied Materials & Interfaces www.acsami.org Research Article

MOG-X themselves are stable (Figure S18). The weight Bo Yuan − State Key Laboratory of NBC Protection for
changes after air exposure are presented in Figure 4b. Among Civilian, Beijing 102205, PR China
the samples desorbed for 7 d, Zr-MOG-1, Zr-MOG-4, Zr- Shouxin Zhang − State Key Laboratory of NBC Protection for
MOG-7, and Zr-MOG-10 display a similar retained absolute Civilian, Beijing 102205, PR China
amount adsorbed. The removed percentage of the adsorbed Guang Yang − State Key Laboratory of NBC Protection for
species in the case of Zr-MOG-10 is evidently lower than that Civilian, Beijing 102205, PR China
of Zr-MOG-1, Zr-MOG-4, and Zr-MOG-7. Thus, for the Complete contact information is available at:
protection against the CWA vapors, Zr-MOG-10 can be https://pubs.acs.org/10.1021/acsami.2c02401
considered as the most competitive protection media for real
applications. Notes
The authors declare no competing financial interest.
4. CONCLUSIONS
A series of granular Zr-MOG-X xerogels with variable
nanoparticle size and defect density are synthesized with
dramatically enhanced surface area and pore volume by tuning
■ ACKNOWLEDGMENTS
We acknowledge the Scientific Compass (https://www.
the amount of concentrated hydrochloric acid and investigated shiyanjia.com/) and the Analytical Instrumentation Center of
for the destruction of the sulfur mustard simulant, 2-CEES. All Peking University for their technical support in character-
the Zr-MOG-X xerogels show outstanding performance for 2- ization. This work was financially supported by the National
CEES degradation with the half-life less than 8.0 min. This is Natural Science Foundation of China (no. 22075319) and the
the highest value reported to date for MOF under pure-phase Huxiang Young Talent Project (2020RC3033).
reaction conditions. The mechanism of decontamination is
that the amino group on the linkers in UiO-66-NH2 MOGs
undergoes a substitution reaction with 2-CEES. The granular
■ REFERENCES
(1) Schwenk, M. Chemical Warfare Agents. Classes and Targets.
Zr-MOG-X xerogels also demonstrate good stability and Toxicol. Lett. 2018, 293, 253−263.
reusability. Static adsorption and desorption tests demonstrate (2) Xu, H.; Gao, Z.; Wang, P.; Xu, B.; Zhang, Y.; Long, L.; Zong, C.;
that Zr-MOG-X xerogels have both a high WU value of up to Guo, L.; Jiang, W.; Ye, Q.; Wang, L.; Xie, J. Biological Effects of
802 mg/g and low desorption capacity with only 28 wt %. Adipocytes in Sulfur Mustard Induced Toxicity. Toxicology 2018, 393,
140−149.
Accordingly, we conclude that granular Zr-MOG-X xerogels (3) Panahi, Y.; Fattahi, A.; Zarei, F.; Ghasemzadeh, N.;
can effectively overcome the issues of Zr-MOF powders and Mohammadpoor, A.; Abroon, S.; Nojadeh, J. N.; Khojastefard, M.;
demonstrate a remarkable prospect in future CWA military Akbarzadeh, A.; Ghasemnejad, T. Next-Generation Sequencing
applications. Approaches for the Study of Genome and Epigenome Toxicity


*
ASSOCIATED CONTENT
sı Supporting Information
Induced by Sulfur Mustard. Arch. Toxicol. 2018, 92, 3443−3457.
(4) Sezigen, S.; Ivelik, K.; Ortatatli, M.; Almacioglu, M.;
Demirkasimoglu, M.; Eyison, R. K.; Kunak, Z. I.; Kenar, L. Victims
of Chemical Terrorism, a Family of Four Who Were Exposed to
The Supporting Information is available free of charge at Sulfur Mustard. Toxicol. Lett. 2019, 303, 9−15.
https://pubs.acs.org/doi/10.1021/acsami.2c02401. (5) Fuchs, A.; Giuliano, E. A.; Sinha, N. R.; Mohan, R. R. Ocular
1 Toxicity of Mustard Gas: A Concise Review. Toxicol. Lett. 2021, 343,
H NMR analysis of Zr-MOG-X; missing linker
evaluated from TGA curves; porosity data of the 21−27.
defected Zr-MOG-X; pore size distributions; relation- (6) Bobbitt, N. S.; Mendonca, M. L.; Howarth, A. J.; Islamoglu, T.;
ship diagram of the half-life with the sizes of the MOF Hupp, J. T.; Farha, O. K.; Snurr, R. Q. Metal-Organic Frameworks for
the Removal of Toxic Industrial Chemicals and Chemical Warfare
catalysts and their defect degrees; characterization and Agents. Chem. Soc. Rev. 2017, 46, 3357−3385.
the 2-CEES removal efficiency of Zr-MOG-0/4 and (7) Islamoglu, T.; Chen, Z.; Wasson, M. C.; Buru, C. T.; Kirlikovali,
their milled powder; N2 adsorption−desorption iso- K. O.; Afrin, U.; Mian, M. R.; Farha, O. K. Metal-Organic
therms and PXRD patterns of Zr-MOG-X after 2-CEES Frameworks against Toxic Chemicals. Chem. Rev. 2020, 120, 8130−
degradation; and GC−MS analysis of acetonitrile 8160.
extraction of 2-CEES degraded by Zr-MOG-4 (PDF) (8) Kirlikovali, K. O.; Chen, Z.; Islamoglu, T.; Hupp, J. T.; Farha, O.
K. Zirconium-based Metal-Organic Frameworks for the Catalytic

■ AUTHOR INFORMATION
Corresponding Authors
Hydrolysis of Organophosphorus Nerve Agents. ACS Appl. Mater.
Interfaces 2020, 12, 14702−14720.
(9) Gil-San-Millan, R.; López-Maya, E.; Hall, M.; Padial, N. M.;
Lin Lu − State Key Laboratory of NBC Protection for Civilian, Peterson, G. W.; DeCoste, J. B.; Rodríguez-Albelo, L. M.; Oltra, J. E.;
Beijing 102205, PR China; Email: 18911066102@189.cn Barea, E.; Navarro, J. A. R. Chemical Warfare Agents Detoxification
Heguo Li − State Key Laboratory of NBC Protection for Properties of Zirconium Metal-Organic Frameworks by Synergistic
Civilian, Beijing 102205, PR China; Email: liheguo1972@ Incorporation of Nucleophilic and Basic Sites. ACS Appl. Mater.
Interfaces 2017, 9, 23967−23973.
126.com
(10) López-Maya, E.; Montoro, C.; Rodríguez-Albelo, L. M.; Aznar
Cheng-an Tao − College of Liberal Arts and Science, National Cervantes, S. D.; Lozano-Pérez, A. A.; Cenís, J. L.; Barea, E.; Navarro,
University of Defense Technology, Changsha 410073, China; J. A. R. Textile/Metal-Organic-Framework Composites as Self-
orcid.org/0000-0003-4720-1064; Email: taochengan@ Detoxifying Filters for Chemical-Warfare Agents. Angew. Chem., Int.
nudt.edu.cn Ed. 2015, 54, 6790−6794.
(11) Ryu, S.-Y.; Chung, J. W.; Kwak, S.-Y. Tunable Multilayer
Authors Assemblies of Nanofibrous Composite Mats as Permeable Protective
Chuan Zhou − State Key Laboratory of NBC Protection for Materials against Chemical Warfare Agents. RSC Adv. 2017, 7, 9964−
Civilian, Beijing 102205, PR China 9974.

23389 https://doi.org/10.1021/acsami.2c02401
ACS Appl. Mater. Interfaces 2022, 14, 23383−23391
ACS Applied Materials & Interfaces www.acsami.org Research Article

(12) Peterson, G. W.; Mahle, J. J.; Tovar, T. M.; Epps, T. H., III (30) Zheng, X.; Zhang, H.; Rehman, S.; Zhang, P. Energy-Efficient
Bent-But-Not-Broken: Reactive Metal-Organic Framework Compo- Capture of Volatile Organic Compounds from Humid Air by
sites from Elastomeric Phase-Inverted Polymers. Adv. Funct. Mater. Granular Metal Organic Gel. J. Hazard. Mater. 2021, 411, 125057−
2020, 30, 2005517−2005524. 125067.
(13) Liu, B.; Vikrant, K.; Kim, K.-H.; Kumar, V.; Kailasa, S. K. (31) Zheng, X.; Wu, Z.; Yang, J.; Rehman, S.; Cao, R.; Zhang, P.
Critical Role of Water Stability in Metal-Organic Frameworks and Metal-Organic Gel Derived N-Doped Granular Carbon: Remarkable
Advanced Modification Strategies for the Extension of Their Toluene Uptake and Rapid Regeneration. ACS Appl. Mater. Interfaces
Applicability. Environ. Sci.: Nano 2020, 7, 1319−1347. 2021, 13, 17543−17553.
(14) Mondal, S. S.; Holdt, H.-J. Breaking Down Chemical Weapons (32) Bueken, B.; Van Velthoven, N.; Willhammar, T.; Stassin, T.;
by Metal-Organic Frameworks. Angew. Chem., Int. Ed. 2016, 55, 42− Stassen, I.; Keen, D. A.; Baron, G. V.; Denayer, J. F. M.; Ameloot, R.;
44. Bals, S.; De Vos, D.; Bennett, T. D. Gel-Based Morphological Design
(15) Buzek, D.; Demel, J.; Lang, K. Zirconium Metal-Organic of Zirconium Metal-Organic Frameworks. Chem. Sci. 2017, 8, 3939−
Framework UiO-66: Stability in an Aqueous Environment and its 3948.
Relevance for Organophosphate Degradation. Inorg. Chem. 2018, 57, (33) Santos-Lorenzo, J.; San José-Velado, R.; Albo, J.; Beobide, G.;
14290−14297. Castaño, P.; Castillo, O.; Luque, A.; Pérez-Yáñez, S. A Straightforward
(16) Jia, M.; Feng, Y.; Qiu, J.; Yao, J. Advances in the Synthesis and Route to Obtain Zirconium Based Metal-Organic Gels. Microporous
Functionalization of UiO-66 and its Applications in Membrane Mesoporous Mater. 2019, 284, 128−132.
Separation. Chem. Ind. Eng. Prog. 2018, 37, 3471−3483. (34) Zheng, X.; He, W.; Rehman, S.; Zhang, P. Facile Synthesis of
(17) Katz, M. J.; Moon, S.-Y.; Mondloch, J. E.; Beyzavi, M. H.; Hydrophobic Metal-Organic Gels for Volatile Organic Compound
Stephenson, C. J.; Hupp, J. T.; Farha, O. K. Exploiting parameter Capture. ACS Appl. Mater. Interfaces 2020, 12, 41359−41367.
space in MOFs: a 20-fold enhancement of phosphate-ester hydrolysis (35) Zhou, C.; Zhang, S.; Pan, H.; Yang, G.; Wang, L.; Tao, C.-a.; Li,
with UiO-66-NH2. Chem. Sci. 2015, 6, 2286−2291. H. Synthesis of Macroscopic Monolithic Metal-Organic Gels for
(18) De Koning, M. C.; Van Grol, M.; Breijaert, T. Degradation of Ultra-Fast Destruction of Chemical Warfare Agents. RSC Adv. 2021,
Paraoxon and the Chemical Warfare Agents VX, Tabun, and Soman 11, 22125−22130.
by the Metal-Organic Frameworks UiO-66-NH2, MOF-808, NU- (36) Katz, M. J.; Brown, Z. J.; Colón, Y. J.; Siu, P. W.; Scheidt, K. A.;
1000, and PCN-777. Inorg. Chem. 2017, 56, 11804−11809. Snurr, R. Q.; Hupp, J. T.; Farha, O. K. A Facile Synthesis of UiO-66,
(19) Ryu, S. G.; Kim, M.-K.; Park, M.; Jang, S. O.; Kim, S. H.; Jung, UiO-67 and Their Derivatives. Chem. Commun. 2013, 49, 9449−
H. Availability of Zr-based MOFs for the Degradation of Nerve 9451.
Agents in All Humidity Conditions. Microporous Mesoporous Mater. (37) Zhao, J.; Chen, R.; Huang, J.; Wang, F.; Tao, C.-A.; Wang, J.
2019, 274, 9−16. Facile Synthesis of Metal-Organic Layers with High Catalytic
(20) Zhang, X.; Sun, Y.; Liu, Y.; Zhai, Z.; Guo, S.; Peng, L.; Qin, Y.; Performance toward Detoxification of a Chemical Warfare Agent
Li, C. UiO-66-NH2 Fabrics: Role of Trifluoroacetic Acid as a Simulant. ACS Appl. Mater. Interfaces 2021, 13, 40863−40871.
Modulator on MOF Uniform Coating on Electrospun Nanofibers and (38) Shearer, G. C.; Chavan, S.; Bordiga, S.; Svelle, S.; Olsbye, U.;
Efficient Decontamination of Chemical Warfare Agent Simulants. Lillerud, K. P. Defect Engineering: Tuning the Porosity and
ACS Appl. Mater. Interfaces 2021, 13, 39976−39984. Composition of the Metal-Organic Framework UiO-66 Via
(21) Yao, Y.; Wang, C.; Na, J.; Hossain, M. S. A.; Yan, X.; Zhang, H.; Modulated Synthesis. Chem. Mater. 2016, 28, 3749−3761.
Amin, M. A.; Qi, J.; Yamauchi, Y.; Li, J. Macroscopic MOF (39) Zhang, X.; Yang, Y.; Lv, X.; Wang, Y.; Liu, N.; Chen, D.; Cui, L.
Architectures: Effective Strategies for Practical Application in Water Adsorption/Desorption Kinetics and Breakthrough of Gaseous
Treatment. Small 2021, 18. DOI: 10.1002/smll.202104387. Toluene for Modified Microporous-Mesoporous UiO-66 Metal
(22) Rego, R. M.; Kuriya, G.; Kurkuri, M. D.; Kigga, M. MOF Based Organic Framework. J. Hazard. Mater. 2019, 366, 140−150.
Engineered Materials in Water Remediation: Recent Trends. J. (40) Shi, X.; Zhang, X.; Bi, F.; Zheng, Z.; Sheng, L.; Xu, J.; Wang, Z.;
Hazard. Mater. 2021, 403, 123605−123634. Yang, Y. Effective toluene adsorption over defective UiO-66-NH2: An
(23) Lorignon, F.; Gossard, A.; Carboni, M. Hierarchically Porous experimental and computational exploration. J. Mol. Liq. 2020, 316,
Monolithic MOFs: An Ongoing Challenge for Industrial-Scale 113812−113824.
Effluent Treatment. Chem. Eng. J. 2020, 393, 124765−124778. (41) Song, L.; Zhao, T.; Yang, D.; Wang, X.; Hao, X.; Liu, Y.; Zhang,
(24) Ryu, U.; Jee, S.; Rao, P. C.; Shin, J.; Ko, C.; Yoon, M.; Park, K. S.; Yu, Z.-Z. Photothermal graphene/UiO-66-NH2 fabrics for
S.; Choi, K. M. Recent Advances in Process Engineering and ultrafast catalytic degradation of chemical warfare agent simulants. J.
Upcoming Applications of Metal-Organic Frameworks. Coord. Chem. Hazard. Mater. 2020, 393, 122332−122343.
Rev. 2021, 426, 213544−213616. (42) Giannakoudakis, D. A.; Bandosz, T. J. Defectous UiO-66 MOF
(25) Li, L.; Xiang, S.; Cao, S.; Zhang, J.; Ouyang, G.; Chen, L.; Su, Nanocomposites as Reactive Media of Superior Protection against
C.-Y. A synthetic route to ultralight hierarchically micro/mesoporous Toxic Vapors. ACS Appl. Mater. Interfaces 2020, 12, 14678−14689.
Al(III)-carboxylate metal-organic aerogels. Nat. Commun. 2013, 4, (43) Oheix, E.; Gravel, E.; Doris, E. Catalytic Processes for the
1774−1782. Neutralization of Sulfur Mustard. Chem.Eur. J. 2021, 27, 54−68.
(26) Hou, J.; Sapnik, A. F.; Bennett, T. D. Metal-Organic (44) Stassen, I.; Bueken, B.; Reinsch, H.; Oudenhoven, J. F. M.;
Framework Gels and Monoliths. Chem. Sci. 2020, 11, 310−323. Wouters, D.; Hajek, J.; Van Speybroeck, V.; Stock, N.; Vereecken, P.
(27) Tian, T.; Zeng, Z.; Vulpe, D.; Casco, M. E.; Divitini, G.; M.; Van Schaijk, R.; De Vos, D.; Ameloot, R. Towards metal-organic
Midgley, P. A.; Silvestre-Albero, J.; Tan, J.-C.; Moghadam, P. Z.; framework based field effect chemical sensors: UiO-66-NH2 for nerve
Fairen-Jimenez, D. A Sol-Gel Monolithic Metal-Organic Framework agent detection. Chem. Sci. 2016, 7, 5827−5832.
with Enhanced Methane Uptake. Nat. Mater. 2018, 17, 174−179. (45) Wang, G.; Sharp, C.; Plonka, A. M.; Wang, Q.; Frenkel, A. I.;
(28) Vilela, S. M. F.; Salcedo-Abraira, P.; Micheron, L.; Solla, E. L.; Guo, W.; Hill, C.; Smith, C.; Kollar, J.; Troya, D.; Morris, J. R.
Yot, P. G.; Horcajada, P. A Robust Monolithic Metal-Organic Mechanism and Kinetics for Reaction of the Chemical Warfare Agent
Framework with Hierarchical Porosity. Chem. Commun. 2018, 54, Simulant, DMMP(g), with Zirconium(IV) MOFs: An Ultrahigh-
13088−13091. Vacuum and DFT Study. J. Phys. Chem. C 2017, 121, 11261−11272.
(29) Connolly, B. M.; Aragones-Anglada, M.; Gandara-Loe, J.; (46) Katz, M. J.; Mondloch, J. E.; Totten, R. K.; Park, J. K.; Nguyen,
Danaf, N. A.; Lamb, D. C.; Mehta, J. P.; Vulpe, D.; Wuttke, S.; S. T.; Farha, O. K.; Hupp, J. T. Simple and Compelling Biomimetic
Silvestre-Albero, J.; Moghadam, P. Z.; Wheatley, A. E. H.; Fairen- Metal-Organic Framework Catalyst for the Degradation of Nerve
Jimenez, D. Tuning Porosity in Macroscopic Monolithic Metal- Agent Simulants. Angew. Chem., Int. Ed. 2014, 53, 497−501.
Organic Frameworks for Exceptional Natural Gas Storage. Nat. (47) Cho, K. Y.; Seo, J. Y.; Kim, H.-J.; Pai, S. J.; Do, X. H.; Yoon, H.
Commun. 2019, 10, 2345−2355. G.; Hwang, S. S.; Han, S. S.; Baek, K.-Y. Facile Control of Defect Site

23390 https://doi.org/10.1021/acsami.2c02401
ACS Appl. Mater. Interfaces 2022, 14, 23383−23391
ACS Applied Materials & Interfaces www.acsami.org Research Article

Density and Particle Size of UiO-66 for Enhanced Hydrolysis Rates: Treatment on the Detoxification of Mustard Gas. Appl. Catal., B
Insights into Feasibility of Zr(IV)-Based Metal-Organic Framework 2018, 226, 429−440.
(MOF) Catalysts. Appl. Catal., B 2019, 245, 635−647. (64) Giannakoudakis, D. A.; Mitchell, J. K.; Bandosz, T. J. Reactive
(48) Li, P.; Klet, R. C.; Moon, S.-Y.; Wang, T. C.; Deria, P.; Peters, Adsorption of Mustard Gas Surrogate on Zirconium (Hydr)oxide/
A. W.; Klahr, B. M.; Park, H.-J.; Al-Juaid, S. S.; Hupp, J. T.; Farha, O. Graphite Oxide Composites: The Role of Surface and Chemical
K. Synthesis of Nanocrystals of Zr-Based Metal-Organic Frameworks Features. J. Mater. Chem. A 2016, 4, 1008−1019.
with Csq-Net: Significant Enhancement in the Degradation of a (65) Giannakoudakis, D. A.; Barczak, M.; Florent, M.; Bandosz, T. J.
Nerve Agent Simulant. Chem. Commun. 2015, 51, 10925−10928. Analysis of Interactions of Mustard Gas Surrogate Vapors with Porous
(49) Zhang, Z.; Tao, C.-A.; Zhao, J.; Wang, F.; Huang, J.; Wang, J. Carbon Textiles. Chem. Eng. J. 2019, 362, 758−766.
Microwave-Assisted Solvothermal Synthesis of UiO-66-NH2 and Its (66) Ploskonka, A. M.; DeCoste, J. B. Tailoring the Adsorption and
Catalytic Performance toward the Hydrolysis of a Nerve Agent Reaction Chemistry of the Metal-Organic Frameworks UiO-66, UiO-
Simulant. Catalysts 2020, 10, 1086−1096. 66-NH2, and HKUST-1 via the Incorporation of Molecular Guests.
(50) Wagner, G. W.; Bartram, P. W.; Koper, O.; Klabunde, K. J. ACS Appl. Mater. Interfaces 2017, 9, 21579−21585.
Reactions of VX, GD, and HD with Nanosize MgO. J. Phys. Chem. B (67) Song, Y.; Chau, J.; Sirkar, K. K.; Peterson, G. W.; Beuscher, U.
1999, 103, 3225−3228. Membrane-Supported Metal Organic Framework Based Nanopacked
(51) Wagner, G. W.; Koper, O. B.; Lucas, E.; Decker, S.; Klabunde, Bed for Protection against Toxic Vapors and Gases. Sep. Purif.
K. J. Reactions of VX, GD, and HD with Nanosize CaO: Autocatalytic Technol. 2020, 251, 117406−117416.
Dehydrohalogenation of HD. J. Phys. Chem. B 2000, 104, 5118−5123.
(52) Wagner, G. W.; Procell, L. R.; O’Connor, R. J.; Munavalli, S.;
Carnes, C. L.; Kapoor, P. N.; Klabunde, K. J. Reactions of VX, GB,
GD, and HD with Nanosize Al2O3. Formation of Aluminophosph-
onates. J. Am. Chem. Soc. 2001, 123, 1636−1644.
(53) Prasad, G. K.; Singh, B.; Ganesan, K.; Batra, A.; Kumeria, T.;
Gutch, P. K.; Vijayaraghavan, R. Modified Titania Nanotubes for
Decontamination of Sulphur Mustard. J. Hazard. Mater. 2009, 167,
1192−1197.
(54) Prasad, G. K.; Mahato, T. H.; Pandey, P.; Singh, B.;
Suryanarayana, M. V. S.; Saxena, A.; Shekhar, K. Reactive sorbent
based on manganese oxide nanotubes and nanosheets for the
decontamination of 2-chloro-ethyl ethyl sulphide. Microporous
Mesoporous Mater. 2007, 106, 256−261.
(55) Mahato, T. H.; Prasad, G. K.; Singh, B.; Srivastava, A. R.;
Ganesan, K.; Acharya, J.; Vijayaraghavan, R. Reactions of sulphur
mustard and sarin on V1.02O2.98 nanotubes. J. Hazard. Mater. 2009,
166, 1545−1549.
(56) Sadeghi, M.; Ghaedi, H.; Yekta, S.; Babanezhad, E.
Decontamination of Toxic Chemical Warfare Sulfur Mustard and
Recommended by ACS
Nerve Agent Simulants by NiO NPs/Ag-Clinoptilolite Zeolite
Composite Adsorbent. J. Environ. Chem. Eng. 2016, 4, 2990−3000. Rational Design of a Zr-MOF@Curli-Polyelectrolyte Hybrid
(57) Sharma, A.; Singh, B.; Saxena, A. Polyoxometalate Impregnated Membrane toward Efficient Chemical Protection, Moisture
Carbon Systems for the in Situ Degradation of Sulphur Mustard. Permeation, and Catalytic Detoxification
Carbon 2009, 47, 1911−1915. Jing Liu, Jinyi Zhong, et al.
(58) Saxena, A.; Singh, B.; Srivastava, A. K.; Suryanarayana, M. V. S.; NOVEMBER 16, 2022
Ganesan, K.; Vijayaraghavan, R.; Dwivedi, K. K. Al2O3 nanoparticles ACS APPLIED MATERIALS & INTERFACES READ
with and without polyoxometalates as reactive sorbents for the
removal of sulphur mustard. Microporous Mesoporous Mater. 2008, Washable and Reusable Zr-Metal–Organic Framework
115, 364−375. Nanostructure/Polyacrylonitrile Fibrous Mats for Catalytic
(59) Roy, A.; Srivastava, A. K.; Singh, B.; Shah, D.; Mahato, T. H.; Degradation of Real Chemical Warfare Agents
Srivastava, A. Kinetics of Degradation of Sulfur Mustard and Sarin
Youngsang Ko, U-Hwang Lee, et al.
Simulants on HKUST-1 Metal Organic Framework. Dalton Trans.
JULY 11, 2022
2012, 41, 12346−12348. ACS APPLIED NANO MATERIALS READ
(60) Li, Y.; Gao, Q.; Zhou, Y.; Zhang, L.; Zhong, Y.; Ying, Y.; Zhang,
M.; Liu, Y.; Wang, Y. Significant Enhancement in Hydrolytic
Zirconium-Based MOFs and Their Biodegradable Polymer
Degradation of Sulfur Mustard Promoted by Silver Nanoparticles in
Composites for Controlled and Sustainable Delivery of
the Ag NPs@HKUST-1 Composites. J. Hazard. Mater. 2018, 358,
Herbicides
113−121.
(61) Kim, H.-H.; Seo, J. Y.; Kim, H.; Jeong, S.; Baek, K.-Y.; Kim, J.; Lila A. M. Mahmoud, Sanjit Nayak, et al.
Min, S.; Kim, S. H.; Jeong, K. Decomposition of the Simulant 2- JULY 29, 2022
ACS APPLIED BIO MATERIALS READ
Chloroethyl Ethyl Sulfide Blister Agent under Ambient Conditions
Using Metal-Organic Frameworks. ACS Appl. Mater. Interfaces 2021,
13, 3782−3792. Enhancing Interfacial and Electromagnetic Interference
(62) Giannakoudakis, D. A.; Pearsall, F.; Florent, M.; Lombardi, J.; Shielding Properties of Carbon Fiber Composites via the
O’Brien, S.; Bandosz, T. J. Barium Titanate Perovskite Nanoparticles Hierarchical Assembly of the MWNT/MOF Interphase
as a Photoreactive Medium for Chemical Warfare Agent Detox- Qingzhong Li, Xiaobiao Zuo, et al.
ification. J. Colloid Interface Sci. 2018, 531, 233−244. NOVEMBER 09, 2022
(63) Giannakoudakis, D. A.; Florent, M.; Wallace, R.; Secor, J.; LANGMUIR READ
Karwacki, C.; Bandosz, T. J. Zinc Peroxide Nanoparticles: Surface,
Chemical and Optical Properties and the Effect of Thermal Get More Suggestions >

23391 https://doi.org/10.1021/acsami.2c02401
ACS Appl. Mater. Interfaces 2022, 14, 23383−23391

You might also like