You are on page 1of 9

www.acsami.

org Research Article

Effective Degradation of Novichok Nerve Agents by the Zirconium


Metal−Organic Framework MOF-808
Martijn C. de Koning,*,§ Carla Vieira Soares,§ Marco van Grol, Rowdy P. T. Bross,
and Guillaume Maurin*
Cite This: ACS Appl. Mater. Interfaces 2022, 14, 9222−9230 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.
Downloaded via UNIV FED DE JUIZ DE FORA on December 12, 2022 at 12:56:23 (UTC).

ABSTRACT: Novichoks are a novel class of nerve agents (also referred to as


the A-series) that were employed in several poisonings over the last few years.
This calls for the development of novel countermeasures that can be applied
in protective concepts (e.g., protective clothing) or in decontamination
methods. The Zr metal−organic framework MOF-808 has recently emerged
as a promising catalyst in the hydrolysis of the V- and G-series of nerve agents
as well as their simulants. In this paper, we report a detailed study of the
degradation of three Novichok agents by MOF-808 in buffers with varying
pH. MOF-808 is revealed to be a highly efficient and regenerable catalyst for
Novichok agent hydrolysis under basic conditions. In contrast to the V- and
G-series of agents, degradation of Novichoks is demonstrated to proceed in
two consecutive hydrolysis steps. Initial extremely rapid P−F bond breaking is followed by MOF-catalyzed removal of the amidine
group from the intermediate product. The intermediate thus acted as a competitive substrate that was rate-determining for the whole
two-step degradation route. Under acidic conditions, the amidine group in Novichok A-230 is more rapidly hydrolyzed than the P−
F bond, giving rise to another moderately toxic intermediate. This intermediate could in turn be efficiently hydrolyzed by MOF-808
under basic conditions. These experimental observations were corroborated by density functional theory calculations to shed light on
molecular mechanisms.
KEYWORDS: Novichoks, MOF-808, DFT, two-step hydrolysis mechanism, degradation

■ INTRODUCTION
Novichok is a subclass of organophosphorus (OP) nerve
member states of the Chemical Weapons Convention (CWC)
to amend for the first time since its ratification in 1997, the
agents, also referred to as the A-series of nerve agents, with a Schedule 1 chemicals list with new agents, including those of
the Novichok series.14
structure similar to the well-described G-series (e.g., sarin,
This motivated further investigations of the chemical and
soman, and tabun) and V-series of nerve agents (e.g., VX and
biological aspects of Novichok agents. Since the first confirmed
VR). These agents are extremely poisonous and exert their
Novichok attack in 2018, a number of reports were edited
toxicity through covalent inhibition of the enzyme acetylcho-
summarizing information on Novichok compounds (almost
linesterase, thereby arresting the hydrolysis of the neuro-
exclusively theoretical data) including possible molecular
transmitter acetylcholine (AChE) in the synapses. This results
structures, physico-chemical, spectroscopic, and toxicological
in overstimulation of AChE receptors and the development of
properties and probable degradation pathways;4,5,15−26 how-
severe clinical signs, for example, convulsions, seizures,
ever, experimental data remain very scarce.27−29 Herein, we
respiratory distress, brain damage, and death.1,2 Although a
unravel for the first time the potential of metal−organic
range of molecular structures belonging to this new class of
frameworks (MOFs) for the catalytic destruction of Novichok
nerve agents have been proposed,3 these structures remained
agents28 (Figure 1). MOFs are highly porous coordination
unverified because of the unavailability of authentic samples
networks built by the assembly of metal ions or clusters with
and the absence of experimental data.4,5 In 2018, an incident
polytopic organic linkers. Over the last decade, MOFs built
occurred in the United Kingdom with a nerve agent that was
later identified as a member of the Novichok family.6,7 In 2020,
another incident involving a Novichok agent occurred in Received: December 15, 2021
Russia.8−10 These events, as well as the use of members of the Accepted: January 26, 2022
G- and V-series of nerve agents in recent history, for example, Published: February 9, 2022
sarin gas use in Syria (2013)11,12 and the use of VX in Malaysia
in 2017,13 illustrate that this type of agents constitute a real
threat to society. The recent Novichok incidents led the 193

© 2022 American Chemical Society https://doi.org/10.1021/acsami.1c24295


9222 ACS Appl. Mater. Interfaces 2022, 14, 9222−9230
ACS Applied Materials & Interfaces www.acsami.org Research Article

OP agents,37,53−55 with recorded half-lives below 1 min for


several real nerve agents and their simulants. MOF-808, built
from tritopic benzene-1,3,5-tricarboxylate (BTC),56 adopts an
spn topology in which 6 of the 12 available coordination sites
on the Zr6-cluster coordinate with 6 BTC linkers (Figure 1),
resulting in a 3D architecture that exhibits 18 Å hexagonal
channels providing sufficient space to accommodate nerve
agent molecules.43 Interestingly, removal of the terminal
formate groups linked to six Zr sites on each cluster results
in an activated form of MOF-808 encompassing exchangeable
hydroxyl- or water ligands (Zr−OH/Zr−OH2) that can act as
Lewis acidic catalytic sites for nerve agent hydrolysis.54,57−59
In this pioneer work, we probed the MOF-808-mediated
hydrolysis of Novichok compounds A-230, A-232, and A-234
(Figure 1) in buffers with varying pH. We demonstrate a rapid
[turnover frequency (TOF) up to 55 min−1] and complete
degradation of Novichok agents via an unusual two-step
hydrolytic pathway. Kinetic models were applied to assess the
catalytic reaction rates. A combination of advanced character-
Figure 1. Molecular structures of the Novichok agents and a ization tools, for example, nuclear magnetic resonance (NMR)
structural representation of MOF-808, showing the Zr6-polyhedra in spectroscopy and high-resolution mass spectrometry (HR-
blue, linkers in gray (hydrogen atoms omitted), and oxygen and MS), enabled the identification of the degradation products as
hydrogen atoms in red and white, respectively. well as the exploration of the regeneration of the catalyst. We
revealed that although the catalyst activity suffers from product
from the Zr6(μ3-O)4(μ3-OH)4 cluster have emerged as inhibition, its activity can be fully restored using a simple
promising catalysts for the hydrolysis of OP nerve agents washing step. These experimental findings were further
from both the G- and V-series and their simulants30−40 and the supported by quantum-calculations at the density functional
influence of MOF topology (connectivity),40 crystal size,41 theory (DFT) level to shed light on the molecular mechanism
pore size,42,43 linker substitution,32,36,44−46 cluster modifica- in play.
tion,46−51 and solvent37,52 on catalytic activity has been
intensively investigated. Among the exhaustive list of Zr-
MOFs, MOF-808 is one of the most effective Zr-MOF
■ RESULTS AND DISCUSSION
Initial Novichok degradation experiments were conducted
catalysts reported to date for the hydrolysis of a wide range of using a commonly employed protocol37,42,60,61 involving the

Figure 2. Degradation profiles of A-230 at pH = 10 using 6.0 mol % MOF-808 (A) and varying the amount of MOF-808 (B). (C) Regeneration of
the catalyst using 0.2 mol % MOF-808.

9223 https://doi.org/10.1021/acsami.1c24295
ACS Appl. Mater. Interfaces 2022, 14, 9222−9230
ACS Applied Materials & Interfaces www.acsami.org Research Article

Table 1. Rate Constants, Half-Lives, and TOFs Determined for the MOF-808-Catalyzed Degradation of Novichok Agents
under Various Conditions
pH/agent MOF-808 (mol %) kobs (× 10−2 min−1)a half-life (min) r2 TOF (min−1)c
pH = 10
A-230 6.0 735 ± 47 0.094 >0.99 122
A-230 1.0 42 ± 3.6 1.7 >0.99 42
A-230 0.5 27 ± 1.5 2.6 >0.99 54
A-230 0.2 11 ± 1 6.4 >0.95 55
A-232 6.0 167 ± 13 0.42 >0.99 28
A-234 6.0 9.5 ± 1.2 7.3 >0.99 1.6
A-234 18 38 ± 4 1.8 >0.98 2.1
pH = 4.5
A-230 → 1c (no MOF) N.A. 0.51 ± 0.02b 14 × 101 >0.99 N.A.
A-230 → 1a (no MOF) N.A. 0.086 ± 0.006b 81 × 101 >0.99 N.A.
a
kobs = kfast of the two-phase exponential decay curve fit in Graphpad Prism 8.0.1., see the Supporting Information for details. bA parallel
exponential decay curve fit was used to obtain these values, see the Supporting Information for details. cTOF was calculated44,52 by dividing the
initial rate (= kfast × [Agent]0 in mM/min) by the catalyst loading (= mol % MOF × [Agent]0 in mM), with [Agent]0 = initial agent concentration
(25 mM in all cases). N.A. = Not Applicable.

addition of 6.0 mol % MOF-808 (particle size: 650 nm) in a MOF-808, while the second (slow) phase represents the
solution of an agent (25 mM) in ethylmorpholine buffer hydrolysis rate of A-230 in competition with 1a. We applied
(NEM, 0.4 M) at pH = 10. The degradation progress was the same method to the experimental data obtained with lower
monitored by consecutive 31P NMR measurements. The amounts of MOF (i.e., 1.0, 0.5, and 0.2 mol %, see Figure 2B).
results obtained with A-230 are shown in Figure 2A. The turning point between the fast phase and the slow phase
While A-230 was stable in NEM buffer pH = 10 for at least was consistently found at 70 ± 2% conversion of A-230, which
60 min, the addition of MOF-808 (6 mol %) resulted in a rapid in all cases coincided with the time at which the highest
degradation of A-230 to about 75% completion within 1 min concentration of 1a (50−60%, see Figure S6) was measured.
(Figure 2A). Another 15 min was required to reach full This supports the assumption that competition between 1a
conversion. This effect could not be ascribed to a pH change, and A-230 was the major cause of the biphasic degradation
which remained the same throughout the experiment. Rather, behavior, although an additional effect of 1b cannot be ruled
this behavior can be explained by a rapid initial hydrolysis of out, vide infra. The fast phase rate constants and corresponding
the P−F bond in A-230, giving rise to the formation of half-lives are provided in Table 1. The calculation of the TOF
hydrolysis product 1a and a subsequent catalytic conversion of of MOF-808 was accomplished by dividing the initial
1a into methylphosphonic acid (1b) through the catalyzed loss degradation rate (in mM/min) by the catalyst concentration
of the amidine moiety. Thus, once 1a reaches a sufficiently (in mM) to provide for a more ready comparison of the
high concentration, it is expected to compete with the catalytic activity under varying conditions. For the degradation
remaining A-230 for catalytic sites in MOF-808, resulting in of A-230 at pH 10, the TOF values were 55, 54, and 42 min−1
slower conversion of A-230. NMR analysis of a sample, for the 0.2, 0.5, and 1.0 mol % reactions, respectively. The
obtained after removal of the MOF by filtration at an early rather high TOF obtained for the 6 mol % reaction (122
time-point, confirmed that hydrolysis of A-230 as well as of 1a min−1) was ascribed to the extremely fast reaction rate at early
was MOF-catalyzed as both compounds remained stable for at time-points. In these cases, the short time needed for filtration,
least 60 min after the MOF removal (Figure S5). The albeit only seconds was not included in the time-point
attenuation of A-230 hydrolysis by 1a is reminiscent of the determination, leading to an overestimation of the reaction
commonly observed inhibition of the catalytic activity of OP rate.
degrading MOFs, which was attributed to the occupancy of the Next, the performance of the catalyst was investigated in
catalytic sites by the anionic hydrolysis product.37,42,62−64 This more detail. Thus, a fresh amount of A-230, equivalent to
phenomenon leads to a progressive deviation from the another 25 mM, was added to the mixture obtained 90 min
expected first-order hydrolysis kinetics model with increasing after reacting 0.2 mol % MOF with A-230 in a pH = 10 buffer.
substrate conversion, and therefore initial rates are often Figure 2C (black curve) shows that the freshly added A-230
reported. Liao et al. recently described62 a comprehensive was degraded significantly slower, which may be attributed to
methodology to more accurately determine the reaction rate MOF degradation over time or occupation of catalytic sites by
constant of the MOF-catalyzed conversion of the nerve agent 1b (catalyst poisoning/inhibition). Remarkably, the catalytic
simulant 4-nitrophenyl dimethylphosphate (DMNP) into activity of MOF-808 is fully restored by a single wash of the
dimethylphosphate (DMP) by NU-1000 and MOF-808, by MOF with fresh buffer, after its isolation by centrifugation, as
implementing the separately determined inhibition constant of shown in Figure 2C (blue curve), suggesting that 1b (being the
DMP onto the catalyst in the kinetic model. However, in the major component present after degradation of A-230 and 1a)
current situation, the pre-determination of the binding/affinity is also capable of catalyst inhibition. Nevertheless, the
constant of the initial product (1a) is problematic as 1a is, in complete regeneration of catalyst activity demonstrates its
contrast to DMNP, in turn consumed. We therefore used a robustness in the degradation of A-230 and confirms the
biphasic exponential decay kinetic model to approximate the reversible nature of catalyst inhibition.
rate constants, assuming that the first (fast) phase is We then explored the degradation of A-232 and A-234
representative of the initial rapid degradation of A-230 by (Figure 3A,B, respectively) employing the same method (6
9224 https://doi.org/10.1021/acsami.1c24295
ACS Appl. Mater. Interfaces 2022, 14, 9222−9230
ACS Applied Materials & Interfaces www.acsami.org Research Article

Figure 3. Degradation profiles of (A) A-232 with 6 mol % MOF-808; (B) A-234 with 6 and 18 mol % MOF-808 both at pH = 10.

Figure 4. (A) Degradation profiles of A-230 at pH = 4.5 in the presence (A) and in the absence (B) of 6 mol % MOF-808. (C) Demonstration of
catalyzed degradation of both 1a and 1c at pH = 10 by MOF-808 (6 mol % with respect to the total phosphorus content).

mol % catalyst in pH = 10 buffer). The reactions proceeded times higher than that of A-232, which can be explained by a
according to the same molecular conversions as found with A- reactivity difference as a result of the change of the P−Me
230. The reaction with A-232 showed a rapid onset of group in A-230 into a deactivating P−OMe group in A-232.
conversion (47% in 1 min) but then required another 27 min The effect of a single methylene group difference between the
to reach completion. A-234 degradation only reached 65% structures of A-232 (P−OMe) and A-234 (P−OEt) on the
conversion within 60 min. Even after increasing the MOF degradation rates of these molecules was remarkable (the TOF
amount to 18 mol %, more than 60 min were required to reach of A-232 was approximately 10 times higher than the TOF of
near complete degradation of A-234 (Figure 3B, 18 mol % A-234).
curve). Similarly, conversion of the respective initial products We next probed the influence of pH on the conversion rates
2a and 3a into the corresponding phosphonic acids 2b and 3b by executing the same reactions in aqueous buffers at pH = 7
was slower (<10% of 2b and <5% of 3b after 60 min with the 6 and pH = 4.5. These pH values were chosen in between the
mol % catalyst). The TOF of A-230 was approximately 1.7 pKa values previously determined65 for the Zr-μ3-OH (3.64),
9225 https://doi.org/10.1021/acsami.1c24295
ACS Appl. Mater. Interfaces 2022, 14, 9222−9230
ACS Applied Materials & Interfaces www.acsami.org Research Article

Figure 5. Illustration of the hydrolysis degradation mechanism of A-230 by MOF-808(Zr). (a) Simulated MEP, (b) reactant (MOF-808 + A-230),
(c) transition state (TS), and (d) product (MOF-808-1a + HF). The most representative separating distances between the molecule and the
inorganic node of MOF-808 involved in the degradation process, F−H(OH) (blue) and P−O(OH) (red), are represented and expressed in Å.

Zr−OH2 (6.22), and Zr−OH (8.23) protons to probe the hydrolysis of the Novichoks A-230, A-232, and A-234 with
reactivity of MOF-808 with different proton topologies. MOF-808, a higher pH is required compared to the hydrolysis
Previous reports showed that Zr MOFs (including MOF- of VX and DMNP by the same MOF.
808) can be effective catalysts at a pH significantly lower than The results of the reactions at pH = 4.5 are provided in
10, which was attributed to the presence of a higher fraction of Figure 4. A-232 and A-234 were not degraded in the absence
Zr−OH2 species in the MOF.37,52 Hydrolysis experiments at or presence of MOF-808 (6 mol %) at pH = 4.5 (Figure S8).
pH = 4.5 are not commonly executed because the low OH− In the presence of the MOF, the more reactive A-230
concentration associated with this pH is expected to give less degraded slowly but completely within 16 h, giving a mixture
effective hydrolysis and/or regeneration of the catalyst, but we of 1a (20%), 1b (6%), and compound 1c (74%) as shown in
reasoned that the unique amidine feature in the Novichok Figure 4A. The latter compound, resulting from P−N
compounds may be protonated and encouraged to act as a hydrolysis in A-230, was not detected at a higher pH. The
leaving group at a lower pH. The predicted pKa of the amidine reaction in the absence of MOF-808 led to the same
group in A-230 is 8.0 and 7.1 for both the amidine groups in degradation rate of A-230, and after complete consumption
A-232 and A-234.66 (16 h), a mixture was formed of 1a (12%) and 1c (88%), as
At pH = 7, the same conversions and similar reactivity illustrated in Figure 4B. Compound 1b was not detected.
difference trends between the agents were observed as at pH = Apparently, the presence of MOF-808 did not significantly
10, but overall, the reactions were much slower. Thus, the accelerate A-230 degradation, but it was not completely
addition of 6 mol % MOF-808 at pH = 7 gave 36 and 17% inactive, judging from the difference in ratio of products in
conversion of A-230 and A-232, respectively, after 60 min Figure 4A,B and the formation of 1b in the presence of MOF.
(Figure S7), while A-234 degradation was very slow under The rate constants for the spontaneous hydrolysis reactions
these conditions (<5% after 60 min). Compound 1a was barely (A-230 → 1a and A-230 → 1c, in the absence of MOF) were
converted into 1b (<3% presence after 60 min) and the determined by fitting a parallel reactions kinetic model (Table
formation of 2b was not detected. This result is in sharp 1).
contrast with previous findings that showed the rapid Finally, we aimed to assess whether MOF-808 can convert
conversion of the nerve agent VX or the nerve agent simulant 1c into 1b at pH = 10. The relevance of this experiment was
DMNP by Zr-MOFs at pH 7,67 indicating that for the efficient prompted by the 2 orders of magnitude higher toxicity of 1c
9226 https://doi.org/10.1021/acsami.1c24295
ACS Appl. Mater. Interfaces 2022, 14, 9222−9230
ACS Applied Materials & Interfaces www.acsami.org Research Article

Figure 6. Illustration of the hydrolysis degradation mechanism of 1a by MOF-808(Zr). (a) Simulated MEP, (b) reactant (MOF-808 + 1a), (c)
transition state (TS), and (d) product (MOF-808-1b + C6H14N2). The main representative separating distances between the molecule and the
inorganic node of MOF-808 involved in the degradation process, P−N (dark blue) and P−O(H2O) (red), are represented and expressed in Å.

[LD50 77.59 mg/kg (mice)]68 compared to 1b [LD50 > 5000 distances of 2.1 and 2.2 Å, respectively (Figure 5C). The
mg/kg (rat)].69 Thus, a fresh mixture of 1a/1c (13:87 mol/ hydrolysis reaction further proceeds with the nucleophilic
mol) was obtained by subjecting A-230 to a pH = 4.5 buffer attack on A-230, resulting in fluoride elimination, where the
for at least 18 h. Next, the pH was adjusted to 10, and the hydrolysis product 1a binds in a bidentate fashion to the
solution was split into two portions. MOF-808 was added to inorganic node (Figure 5D). The resulting MEP, reported in
one of the portions, and both solutions were monitored using Figure 5A, implies an activation energy barrier of 109.7 kJ
31
P NMR for another 7 h (Figure 4C). In the absence of MOF- mol−1. This value is within the same order of magnitude as the
808, 1a and 1c were stable at pH = 10 for at least 7 h,70 while free energy barriers previously reported for the hydrolysis of
both 1a and 1c were converted into 1b in the presence of sarin using MOF-808(Zr) (79.1 kJ mol−1), NU-1000(Zr)
MOF-808, confirming the capability of MOF-808 to catalyze (75.3 kJ mol−1), and UiO-66(Zr) (71.5 kJ mol−1).71 Since
the conversion of 1a into 1b and emphasizing the ability of these Zr-MOFs were previously demonstrated to be effective
MOF-808 to catalyze the hydrolysis of 1c (→ 1b) as well. for the degradation of sarin,37 the similarity obtained in terms
DFT calculations were further performed for A-230 and A- of the predicted energy barrier supports the fact that MOF-808
232 to shed light on the catalytic degradation mechanism of is responsible of the degradation of Novichok observed
Novichok compounds. The first degradation step corresponds experimentally. In addition, the lowest energy obtained for
to the hydrolysis of the P−F bond in A-230 induced by the the product (−74.5 kJ mol−1) indicates that the coordination
nucleophilic addition of the −OH group to the P atom, as of 1a with the inorganic node of MOF-808 leads to the
illustrated in Figure 5, which reports the minimum energy path formation of a very stable complex, consistent with the
(MEP), the snapshots of the reactant, the transition state, and previous calculations reported by Snurr et al.54 as well as with
the product for the corresponding reaction. Figure 5B shows the current experimental results.
the initial binding of A-230 via its PO function to an The second step in the hydrolysis sequence (1a → 1b) was
uncoordinated Zr-site generated by the displacement of a also the subject of DFT calculations. The MEP for the
terminal water molecule. At the transition state, the molecule conversion of MOF-808-1a into MOF-808-1b is reported in
approaches the −OH moiety, with F−H(OH) and P−O(OH) Figure 6. This second step proceeds via a nucleophilic attack of
9227 https://doi.org/10.1021/acsami.1c24295
ACS Appl. Mater. Interfaces 2022, 14, 9222−9230
ACS Applied Materials & Interfaces www.acsami.org Research Article

an extra water molecule on the P atom of 1a (Figure 6B), evaluated to probe the applicability in, for instance, protective
resulting in the loss of the amidine fragment and the formation clothing. The latter issue is more generally applicable to other
of methylphosphonic acid (MPA) (Figure 6C). Beyond the agents as well, and may, for instance, be overcome by
transition states, the final product 1b adopts a bidentate implementing basic polymers.43,64,74,75
coordination with the inorganic node, where the separating
distance between the amidine molecule and the MPA increases
from 2.2 to 3.6 Å (dark dashed blue line) (Figure 6B,C). The

*
ASSOCIATED CONTENT
sı Supporting Information
resulting much higher activation energy (196.1 kJ mol−1) The Supporting Information is available free of charge at
compared to the one calculated for the A-230 → 1a conversion https://pubs.acs.org/doi/10.1021/acsami.1c24295.
(109.7 kJ mol−1) clearly highlights that the catalytic conversion
of 1a to 1b is the rate-limiting step reaction of the overall Chemicals and instrumentation details; experimental
degradation mechanism of Novichok A-230. MOF-808-1a has details on synthesis; degradation experiments and DFT
the most stable configuration along the entire MEP, which calculations; characterization data of MOF-808 (PXRD,
explains the experimentally observed inhibition of the catalytic N2 isotherms, DLS, and NMR spectra); additional
conversion of A-230 at high 1a concentrations. degradation curves; kinetic models’ description; HR-MS
The degradation mechanism of A-232, detailed in Figure S9 data of the Novichok compounds and their degradation
and in Figure S10, follows the same trend as found for A-230. products; and DFT-obtained MEPs and snapshots of the
Finding similar MEP energy barriers, this calculation supported conversions A-232 → 2a and 2a → 2b (PDF)
the formation of a highly stable product MOF-808-2a, which
coordinates with the inorganic node in a bidentate manner and
which corroborates the inhibition of degradation of the
■ AUTHOR INFORMATION
Corresponding Authors
Novichok compounds and the slow conversion of 2a into 2b. Martijn C. de Koning − TNO Defense, Safety and Security,

■ CONCLUSIONS
We demonstrated the capability of Zr MOF-808 for an
Rijswijk 2288GJ, The Netherlands; orcid.org/0000-
0001-6482-9502; Email: m.dekoning@tno.nl
Guillaume Maurin − ICGM, Univ. Montpellier, CNRS,
effective degradation of Novichok compounds A-230, A-232, ENSCM, Montpellier 34095, France;
and A-234. While stable at pH = 10, A-230 and A-232 were Email: guillaume.maurin1@umontpellier.fr
degraded extremely rapidly in the presence of MOF-808. The
degradation of these compounds followed a two-step Authors
hydrolysis mechanism in which P−F bond hydrolysis resulted Carla Vieira Soares − ICGM, Univ. Montpellier, CNRS,
in the formation of a second, competitive substrate, which, in ENSCM, Montpellier 34095, France
turn, was catalytically degraded through P−N hydrolysis. The Marco van Grol − TNO Defense, Safety and Security, Rijswijk
competitive intermediate species formed a stable complex with 2288GJ, The Netherlands
MOF-808, which attenuated the degradation rate of the parent Rowdy P. T. Bross − TNO Defense, Safety and Security,
compound and which resulted in bi-phasic degradation curves. Rijswijk 2288GJ, The Netherlands; orcid.org/0000-
Kinetic analysis provided the rate constants, half-lives, and 0002-9667-3036
TOFs for the (early stage) interactions between MOF-808 and Complete contact information is available at:
the Novichok compounds. The initial degradation rates of A- https://pubs.acs.org/10.1021/acsami.1c24295
230 and A-232 were very rapid, with half-lives <30s. A-234
hydrolysis proceeded considerably slower with a half-life of Author Contributions
§
approximately 7.3 min. DFT calculations confirmed the M.C.d.K. and C.V.S contributed equally to this work.
reaction sequences observed experimentally and predicted Notes
the formation of the highly stable intermediate, observed The authors declare no competing financial interest.


experimentally, which acted as the rate-limiting step in the two-
step hydrolysis sequence. Although the hydrolysis products ACKNOWLEDGMENTS
inhibited the catalyst, full recovery of catalytic activity could be
M.C.d.K., M.v.G., and R.B. thank the Dutch MoD for financial
effected with only a single wash of the MOF with a buffer,
support in the CBRN research program V1802 and the
showing the potential of MOF-808 as an effective catalyst for
CPIUM Program. The authors thank Wouter van de Steeg for
detoxifying nerve agents, including Novichoks. However, the
SEM images and Thomas Hartman (thisillustrations.com) for
use of pH = 7 and pH = 4.5 buffers resulted in a significant
the production of the graphical abstract. C.V.S. and G.M. thank
reduction in conversions rates, showing that a high pH (>7) is
the ANR SESAM for funding. The computational work was
required for efficient Novichok agent degradation by MOF-
performed using HPC resources from GENCI-CINES (grant
808. At pH = 4.5, MOF-808 was marginally active, but the low
A0100907613).


pH gave rise to 6 times more rapid P−N bond hydrolysis than
P−F bond hydrolysis in A-230, leading to the formation of the REFERENCES
moderately toxic compound 1c as the major product. The
latter product could be effectively converted into non-toxic 1b (1) Marrs, T. C. Organophosphate Poisoning. Pharmacol. Ther.
1993, 58, 51−66.
by MOF-808 at pH = 10. Overall, the results show that the (2) Sussman, J. L.; Harel, M.; Frolow, F.; Oefner, C.; Goldman, A.;
current examples of the new Novichok nerve agents can be Toker, L.; Silman, I. Atomic Structure of Acetylcholinesterase from
effectively degraded by MOF-808 under basic, aqueous Torpedo Californica: A Prototypic Acetylcholine-Binding Protein.
conditions. These findings encourage further research toward Science 1991, 253, 872−879.
their applicability in (skin) decontamination concepts.72,73 The (3) Marzayanov, V. S. State Secrets: An Insider’s Chronicle of the
performance under buffer-free conditions needs to be Russian Chemical Weapons Program; Outskirts Press Inc, 2009.

9228 https://doi.org/10.1021/acsami.1c24295
ACS Appl. Mater. Interfaces 2022, 14, 9222−9230
ACS Applied Materials & Interfaces www.acsami.org Research Article

(4) Nepovimova, E.; Kuca, K. Chemical Warfare Agent NOV- (A234)”, in Aqueous Solution by Liquid Chromatography-Tandem
ICHOK - Mini-Review of Available Data. Food Chem. Toxicol. 2018, Mass Spectrometry. Molecules 2021, 26, 1059.
121, 343−350. (28) Harvey, S. P.; McMahon, L. R.; Berg, F. J. Hydrolysis and
(5) Kloske, M.; Witkiewicz, Z. Novichoks − The A Group of Enzymatic Degradation of Novichok Nerve Agents. Heliyon 2020, 6,
Organophosphorus Chemical Warfare Agents. Chemosphere 2019, No. e03153.
221, 672−682. (29) Jacquet, P.; Rémy, B.; Bross, R. P. T.; van Grol, M.; Gaucher,
(6) Mark Peplow, M. Nerve Agent Attack Used Novichok Poison. F.; Chabrière, E.; de Koning, M. C.; Daudé, D. Enzymatic
Chem. Eng. News 2018, 96, 3. Decontamination of G-Type, V-Type and Novichok Nerve Agents.
(7) Vale, J. A.; Marrs, T. C.; Maynard, R. L. Novichok: A Murderous Int. J. Mol. Sci. 2021, 22, 8152.
Nerve Agent Attack in the UK. Clin. Toxicol. 2018, 56, 1093−1097. (30) Liu, Y.; Howarth, A. J.; Vermeulen, N. A.; Moon, S.-Y.; Hupp, J.
(8) Steindl, D.; Boehmerle, W.; Körner, R.; Praeger, D.; Haug, M.; T.; Farha, O. K. Catalytic Degradation of Chemical Warfare Agents
Nee, J.; Schreiber, A.; Scheibe, F.; Demin, K.; Jacoby, P.; Tauber, R.; and Their Simulants by Metal-Organic Frameworks. Coord. Chem.
Hartwig, S.; Endres, M.; Eckardt, K.-U. Novichok Nerve Agent Rev. 2017, 346, 101−111.
Poisoning. Lancet 2021, 397, 249−252. (31) Mondloch, J. E.; Katz, M. J.; Isley, W. C.; Ghosh, P.; Liao, P.;
(9) Seibert, S. Statement by the Federal Government on the Navalny Bury, W.; Wagner, G. W.; Hall, M. G.; DeCoste, J. B.; Peterson, G.
Case https://archiv.bundesregierung.de/archiv-de/meta/startseite/ W.; Snurr, R. Q.; Cramer, C. J.; Hupp, J. T.; Farha, O. K. Destruction
statement-by-the-federal-government-on-the-navalny-case-1781882 of Chemical Warfare Agents Using Metal-Organic Frameworks. Nat.
(accessed Nov 15, 2021). Mater. 2015, 14, 512−516.
(10) Eddleston, M.; Chowdhury, F. R. Organophosphorus Poison- (32) Kalaj, M.; Momeni, M. R.; Bentz, K. C.; Barcus, K. S.; Palomba,
ing: The Wet Opioid Toxidrome. Lancet 2021, 397, 175−177. J. M.; Paesani, F.; Cohen, S. M. Halogen Bonding in UiO-66
(11) John, H.; van der Schans, M. J.; Koller, M.; Spruit, H. E. T.; Frameworks Promotes Superior Chemical Warfare Agent Simulant
Worek, F.; Thiermann, H.; Noort, D. Fatal Sarin Poisoning in Syria Degradation. Chem. Commun. 2019, 55, 3481−3484.
2013: Forensic Verification within an International Laboratory (33) Bobbitt, N. S.; Mendonca, M. L.; Howarth, A. J.; Islamoglu, T.;
Network. Forensic Toxicol. 2018, 36, 61−71. Hupp, J. T.; Farha, O. K.; Snurr, R. Q. Metal−Organic Frameworks
(12) Dolgin, E. Syrian Gas Attack Reinforces Need for Better Anti- for the Removal of Toxic Industrial Chemicals and Chemical Warfare
Sarin Drugs. Nat. Med. 2013, 19, 1194−1195. Agents. Chem. Soc. Rev. 2017, 46, 3357−3385.
(13) Nakagawa, T.; Tu, A. T. Murders with VX: Aum Shinrikyo in (34) Zhang, Y.; Zhang, X.; Lyu, J.; Otake, K.-i.; Wang, X.; Redfern,
Japan and the Assassination of Kim Jong-Nam in Malaysia. Forensic L. R.; Malliakas, C. D.; Li, Z.; Islamoglu, T.; Wang, B.; Farha, O. K. A
Toxicol. 2018, 36, 542−544. Flexible Metal-Organic Framework with 4-Connected Zr6 Nodes. J.
(14) OPCW. Change to Schedule 1 of the annex on chemical Am. Chem. Soc. 2018, 140, 11179−11183.
weapons convention https://www.opcw.org/sites/default/files/ (35) Liang, H.; Yao, A.; Jiao, X.; Li, C.; Chen, D. Fast and Sustained
documents/2019/11/c24dec05%28e%29.pdf (accessed Nov 15, Degradation of Chemical Warfare Agent Simulants Using Flexible
2021). Self-Supported Metal-Organic Framework Filters. ACS Appl. Mater.
(15) Patocka, J. Novichok Agents-Mysterious Poisonous Substances Interfaces 2018, 10, 20396−20403.
from the Cold War Period. Vojen. Zdrav. Listy 2018, 87, 92−94. (36) Islamoglu, T.; Ortuño, M. A.; Proussaloglou, E.; Howarth, A. J.;
(16) Sajid, H.; Khan, S.; Ayub, K.; Mahmood, T. Effective Vermeulen, N. A.; Atilgan, A.; Asiri, A. M.; Cramer, C. J.; Farha, O. K.
Adsorption of A-Series Chemical Warfare Agents on Graphdiyne Presence versus Proximity: The Role of Pendant Amines in the
Nanoflake: A DFT Study. J. Mol. Model. 2021, 27, 117. Catalytic Hydrolysis of a Nerve Agent Simulant. Angew. Chem., Int.
(17) Yar, M.; Hashmi, M. A.; Khan, A.; Ayub, K. Carbon Nitride 2- Ed. 2018, 57, 1949−1953.
D Surface as a Highly Selective Electrochemical Sensor for V-Series (37) de Koning, M. C.; van Grol, M.; Breijaert, T. Degradation of
Nerve Agents. J. Mol. Liq. 2020, 311, 113357. Paraoxon and the Chemical Warfare Agents VX, Tabun, and Soman
(18) Motlagh, N. M.; Rouhani, M.; Mirjafary, Z. Aminated C20 by the Metal−Organic Frameworks UiO-66-NH2, MOF-808, NU-
Fullerene as a Promising Nanosensor for Detection of A-234 Nerve 1000, and PCN-777. Inorg. Chem. 2017, 56, 11804−11809.
Agent. Comput. Theor. Chem. 2020, 1186, 112907. (38) Chen, Z.; Wang, X.; Noh, H.; Ayoub, G.; Peterson, G. W.;
(19) Imrit, Y. A.; Bhakhoa, H.; Sergeieva, T.; Danés, S.; Savoo, N.; Buru, C. T.; Islamoglu, T.; Farha, O. K. Scalable, Room Temperature,
Elzagheid, M. I.; Rhyman, L.; Andrada, D. M.; Ramasami, P. A and Water-Based Synthesis of Functionalized Zirconium-Based Metal-
theoretical study of the hydrolysis mechanism of A-234; the suspected Organic Frameworks for Toxic Chemical Removal. CrystEngComm
novichok agent in the Skripal attack. RSC Adv. 2020, 10, 27884− 2019, 21, 2409−2415.
27893. (39) Wang, H.; Mahle, J. J.; Tovar, T. M.; Peterson, G. W.; Hall, M.
(20) Carlsen, L. After Salisbury Nerve Agents Revisited. Mol. Inf. G.; Decoste, J. B.; Buchanan, J. H.; Karwacki, C. J. Solid-Phase
2019, 38, No. e1800106. Detoxification of Chemical Warfare Agents Using Zirconium-Based
(21) Jeong, K.; Choi, J. Theoretical Study on the Toxicity of Metal Organic Frameworks and the Moisture Effects: Analyze via
‘Novichok’ Agent Candidates. R. Soc. Open Sci. 2019, 6, 190414. Digestion. ACS Appl. Mater. Interfaces 2019, 11, 21109−21116.
(22) Hussain, N. M.; Sharma, S. C. Novichok: An Overview of the (40) Son, F. A.; Wasson, M. C.; Islamoglu, T.; Chen, Z.; Gong, X.;
World’s Deadliest Nerve Agent. Br. Student Dr. J. 2019, 3, 48. Hanna, S. L.; Lyu, J.; Wang, X.; Idrees, K. B.; Mahle, J. J.; Peterson, G.
(23) Franca, T.; Kitagawa, D.; Cavalcante, S.; da Silva, J.; W.; Farha, O. K.; Farha, O. K. Uncovering the Role of Metal-Organic
Nepovimova, E.; Kuca, K. Novichoks: The Dangerous Fourth Framework Topology on the Capture and Reactivity of Chemical
Generation of Chemical Weapons. Int. J. Mol. Sci. 2019, 20, 1222. Warfare Agents. Chem. Mater. 2020, 32, 4609−4617.
(24) Bhakhoa, H.; Rhyman, L.; Ramasami, P. Theoretical Study of (41) Li, P.; Klet, R. C.; Moon, S.-Y.; Wang, T. C.; Deria, P.; Peters,
the Molecular Aspect of the Suspected Novichok Agent A234 of the A. W.; Klahr, B. M.; Park, H.-J.; Al-Juaid, S. S.; Hupp, J. T.; Farha, O.
Skripal Poisoning. R. Soc. Open Sci. 2019, 6, 181831. K. Synthesis of Nanocrystals of Zr-Based Metal−Organic Frameworks
(25) Lyagin, I.; Efremenko, E. Theoretical Evaluation of Suspected with Csq-Net: Significant Enhancement in the Degradation of a
Enzymatic Hydrolysis of Novichok Agents. Catal. Commun. 2019, Nerve Agent Simulant. Chem. Commun. 2015, 51, 10925−10928.
120, 91−94. (42) Peterson, G. W.; Moon, S.-Y.; Wagner, G. W.; Hall, M. G.;
(26) Chai, P. R.; Hayes, B. D.; Erickson, T. B.; Boyer, E. W. Decoste, J. B.; Hupp, J. T.; Farha, O. K. Tailoring the Pore Size and
Novichok Agents: A Historical, Current, and Toxicological Functionality of UiO-Type Metal-Organic Frameworks for Optimal
Perspective. Toxicol. Commun. 2018, 2, 45−48. Nerve Agent Destruction. Inorg. Chem. 2015, 54, 9684−9686.
(27) Lee, J. Y.; Lim, K. C.; Kim, H. S. Characterization and Study on (43) Chen, Z.; Ma, K.; Mahle, J. J.; Wang, H.; Syed, Z. H.; Atilgan,
Fragmentation Pathways of a Novel Nerve Agent, “Novichok A.; Chen, Y.; Xin, J. H.; Islamoglu, T.; Peterson, G. W.; Farha, O. K.

9229 https://doi.org/10.1021/acsami.1c24295
ACS Appl. Mater. Interfaces 2022, 14, 9222−9230
ACS Applied Materials & Interfaces www.acsami.org Research Article

Integration of Metal-Organic Frameworks on Protective Layers for (59) Soares, C. V.; Leitão, A. A.; Maurin, G. Computational
Destruction of Nerve Agents under Relevant Conditions. J. Am. Chem. Evaluation of the Chemical Warfare Agents Capture Performances of
Soc. 2019, 141, 20016−20021. Robust MOFs. Microporous Mesoporous Mater. 2019, 280, 97−104.
(44) Katz, M. J.; Moon, S.-Y.; Mondloch, J. E.; Beyzavi, M. H.; (60) Mian, M. R.; Islamoglu, T.; Afrin, U.; Goswami, S.; Cao, R.;
Stephenson, C. J.; Hupp, J. T.; Farha, O. K. Exploiting Parameter Kirlikovali, K. O.; Hall, M. G.; Peterson, G. W.; Farha, O. K. Catalytic
Space in MOFs: A 20-Fold Enhancement of Phosphate-Ester Degradation of an Organophosphorus Agent at Zn-OH Sites in a
Hydrolysis with UiO-66-NH2. Chem. Sci. 2015, 6, 2286−2291. Metal-Organic Framework. Chem. Mater. 2020, 32, 6998−7004.
(45) Kalaj, M.; Palomba, J. M.; Bentz, K. C.; Cohen, S. M. Multiple (61) Mondloch, J. E.; Katz, M. J.; Isley, W. C.; Ghosh, P.; Liao, P.;
Functional Groups in UiO-66 Improve Chemical Warfare Agent Bury, W.; Wagner, G. W.; Hall, M. G.; DeCoste, J. B.; Peterson, G.
Simulant Degradation. Chem. Commun. 2019, 55, 5367−5370. W.; Snurr, R. Q.; Cramer, C. J.; Hupp, J. T.; Farha, O. K. Destruction
(46) Gil-San-Millan, R.; López-Maya, E.; Hall, M.; Padial, N. M.; of Chemical Warfare Agents Using Metal-Organic Frameworks. Nat.
Peterson, G. W.; DeCoste, J. B.; Rodríguez-Albelo, L. M.; Oltra, J. E.; Mater. 2015, 14, 512−516.
(62) Liao, Y.; Sheridan, T.; Liu, J.; Farha, O.; Hupp, J. Product
Barea, E.; Navarro, J. A. R. Chemical Warfare Agents Detoxification
Inhibition and the Catalytic Destruction of a Nerve Agent Simulant
Properties of Zirconium Metal−Organic Frameworks by Synergistic
by Zirconium-Based Metal-Organic Frameworks. ACS Appl. Mater.
Incorporation of Nucleophilic and Basic Sites. ACS Appl. Mater. Interfaces 2021, 13, 30565−30575.
Interfaces 2017, 9, 23967−23973. (63) Mondloch, J. E.; Katz, M. J.; Isley, W. C.; Ghosh, P.; Liao, P.;
(47) Gil-San-Millan, R.; López-Maya, E.; Platero-Prats, A. E.; Bury, W.; Wagner, G. W.; Hall, M. G.; Decoste, J. B.; Peterson, G. W.;
Torres-Pérez, V.; Delgado, P.; Augustyniak, A. W.; Kim, M. K.; Lee, Snurr, R. Q.; Cramer, C. J.; Hupp, J. T.; Farha, O. K. Destruction of
H. W.; Ryu, S. G.; Navarro, J. A. R. Magnesium Exchanged Zirconium Chemical Warfare Agents Using Metal-Organic Frameworks. Nat.
Metal-Organic Frameworks with Improved Detoxification Properties Mater. 2015, 14, 512−516.
of Nerve Agents. J. Am. Chem. Soc. 2019, 141, 11801−11805. (64) Moon, S. Y.; Proussaloglou, E.; Peterson, G. W.; DeCoste, J. B.;
(48) Garibay, S. J.; Farha, O. K.; Decoste, J. B. Single-Component Hall, M. G.; Howarth, A. J.; Hupp, J. T.; Farha, O. K. Detoxification of
Frameworks for Heterogeneous Catalytic Hydrolysis of Organo- Chemical Warfare Agents Using a Zr6 Based Metal−Organic
phosphorous Compounds in Pure Water. Chem. Commun. 2019, 55, Framework/Polymer Mixture. Chem.Eur. J. 2016, 22, 14864−
7005−7008. 14868.
(49) Cha, G.-Y.; Chun, H.; Hong, D.-Y.; Kim, J.; Cho, K.-H.; Lee, (65) Klet, R. C.; Liu, Y.; Wang, T. C.; Hupp, J. T.; Farha, O. K.
U.-H.; Chang, J.-S.; Ryu, S. G.; Lee, H. W.; Kim, S.-J.; Han, B.; Evaluation of Brønsted Acidity and Proton Topology in Zr- and Hf-
Hwang, Y. K. Unique Design of Superior Metal-Organic Framework Based Metal−Organic Frameworks Using Potentiometric Acid−Base
for Removal of Toxic Chemicals in Humid Environment via Direct Titration. J. Mater. Chem. A 2016, 4, 1479−1485.
Functionalization of the Metal Nodes. J. Hazard. Mater. 2020, 398, (66) Values Were Obtained from Tier-0 Calculations at the
122857. Cheminformatics Website of the ADME Center of Excellence at
(50) Geravand, E.; Farzaneh, F.; Gil-San-Millan, R.; Carmona, F. J.; USAMRICD, USA. http://Cheminfo.Bhsai.Org (accessed 28-8-
Navarro, J. A. R. Mixed-Metal Cerium/Zirconium MOFs with 2020).
Improved Nerve Agent Detoxification Properties. Inorg. Chem. (67) De Koning, M. C.; Peterson, G. W.; Van Grol, M.; Iordanov, I.;
2020, 59, 16160−16167. McEntee, M. Degradation and Detection of the Nerve Agent VX by a
(51) Islamoglu, T.; Atilgan, A.; Moon, S.-Y.; Peterson, G. W.; Chromophore-Functionalized Zirconium MOF. Chem. Mater. 2019,
Decoste, J. B.; Hall, M.; Hupp, J. T.; Farha, O. K. Cerium(IV) vs 31, 7417−7424.
(68) Sun, M.-C.; Zhang, H.; Chou, T.-C. Enzymatic Synthesis of
Zirconium(IV) Based Metal-Organic Frameworks for Detoxification
Sarin and Soman. Biochem. Pharmacol. 1985, 34, 1843−1845.
of a Nerve Agent. Chem. Mater. 2017, 29, 2672−2675.
(69) Munro, N. B.; Talmage, S. S.; Griffin, G. D.; Waters, L. C.;
(52) Katz, M. J.; Klet, R. C.; Moon, S.-Y.; Mondloch, J. E.; Hupp, J.
Watson, A. P.; King, J. F.; Hauschild, V. The Sources, Fate, and
T.; Farha, O. K. One Step Backward Is Two Steps Forward: Toxicity of Chemical Warfare Agent Degradation Products. Environ.
Enhancing the Hydrolysis Rate of UiO-66 by Decreasing [OH−]. ACS Health Perspect. 1999, 107, 933−974.
Catal. 2015, 5, 4637−4642. (70) Worek, F.; Elsinghorst, P.; Koller, M.; Thiermann, H. Reactions
(53) Moon, S.-Y.; Liu, Y.; Hupp, J. T.; Farha, O. K. Instantaneous of Methylphosphonic Difluoride with Human Acetylcholinesterase
Hydrolysis of Nerve-Agent Simulants with a Six-Connected and Oximes - Possible Therapeutic Implications. Toxicol. Lett. 2014,
Zirconium-Based Metal-Organic Framework. Angew. Chem., Int. Ed. 231, 92−98.
Engl. 2015, 54, 6795−6799. (71) Momeni, M. R.; Cramer, C. J. Dual Role of Water in
(54) Mendonca, M. L.; Ray, D.; Cramer, C. J.; Snurr, R. Q. Heterogeneous Catalytic Hydrolysis of Sarin by Zirconium-Based
Exploring the Effects of Node Topology, Connectivity, and Metal Metal-Organic Frameworks. ACS Appl. Mater. Interfaces 2018, 10,
Identity on the Binding of Nerve Agents and Their Hydrolysis 18435−18439.
Products in Metal-Organic Frameworks. ACS Appl. Mater. Interfaces (72) Larsson, A.; Qvarnström, J.; Lindberg, S.; Wigenstam, E.;
2020, 12, 35657−35675. Ö berg, L.; Afshin Sander, R.; Johansson, S.; Bucht, A.; Thors, L. In
(55) Kirlikovali, K. O.; Chen, Z.; Islamoglu, T.; Hupp, J. T.; Farha, Vitro Human Skin Decontamination Efficacy of MOF-808 in
O. K. Zirconium-Based Metal-Organic Frameworks for the Catalytic Decontamination Lotion Following Exposure to the Nerve Agent
Hydrolysis of Organophosphorus Nerve Agents. ACS Appl. Mater. VX. Toxicol. Lett. 2021, 339, 32−38.
Interfaces 2020, 12, 14702−14720. (73) Kalinovskyy, Y.; Wright, A. J.; Hiscock, J. R.; Watts, T. D.;
(56) Jiang, J.; Gándara, F.; Zhang, Y.-B.; Na, K.; Yaghi, O. M.; Williams, R. L.; Cooper, N. J.; Main, M. J.; Holder, S. J.; Blight, B. A.
Klemperer, W. G. Superacidity in Sulfated Metal-Organic Framework- Swell and Destroy: A Metal-Organic Framework-Containing Polymer
808. J. Am. Chem. Soc. 2014, 136, 12844−12847. Sponge That Immobilizes and Catalytically Degrades Nerve Agents.
(57) Troya, D. Reaction Mechanism of Nerve-Agent Decomposition ACS Appl. Mater. Interfaces 2020, 12, 8634−8641.
with Zr-Based Metal Organic Frameworks. J. Phys. Chem. C 2016, (74) Jabbour, C. R.; Parker, L. A.; Hutter, E. M.; Weckhuysen, B. M.
120, 29312−29323. Chemical Targets to Deactivate Biological and Chemical Toxins
(58) Wang, G.; Sharp, C.; Plonka, A. M.; Wang, Q.; Frenkel, A. I.; Using Surfaces and Fabrics. Nat. Rev. Chem. 2021, 5, 370−387.
Guo, W.; Hill, C.; Smith, C.; Kollar, J.; Troya, D.; Morris, J. R. (75) Chen, Z.; Islamoglu, T.; Farha, O. K. Toward Base
Heterogenization: A Zirconium Metal-Organic Framework/Den-
Mechanism and Kinetics for Reaction of the Chemical Warfare Agent
drimer or Polymer Mixture for Rapid Hydrolysis of a Nerve-Agent
Simulant, DMMP(g), with Zirconium(IV) MOFs: An Ultrahigh-
Simulant. ACS Appl. Nano Mater. 2019, 2, 1005−1008.
Vacuum and DFT Study. J. Phys. Chem. C 2017, 121, 11261−11272.

9230 https://doi.org/10.1021/acsami.1c24295
ACS Appl. Mater. Interfaces 2022, 14, 9222−9230

You might also like