Computational Screening of Roles of Defects and Metal Substitution

May 26, 2019 - (3−5) Owing to the well-known high sensitivity of enzymes to ... is fully embraced by the community, and indeed engineering them has ...
0 downloads 0 Views 5MB Size
Article Cite This: J. Phys. Chem. C 2019, 123, 15157−15165

pubs.acs.org/JPCC

Computational Screening of Roles of Defects and Metal Substitution on Reactivity of Different Single- vs Double-Node Metal−Organic Frameworks for Sarin Decomposition Mohammad R. Momeni*,† and Christopher J. Cramer Department of Chemistry, Minnesota Supercomputing Institute, Chemical Theory Center, University of Minnesota, Minneapolis, Minnesota 55455, United States Downloaded via KEAN UNIV on July 30, 2019 at 13:29:24 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

S Supporting Information *

ABSTRACT: Understanding how different factors affect the electronic properties of metal−organic frameworks (MOFs) is critical to understanding their potential for catalysis and to serve as catalyst supports. In this work, periodic dispersioncorrected quantum mechanical calculations are performed to assess the catalytic activity of different Zr6 vs Zr12 MOFs for the heterogeneous catalytic hydrolysis of the chemical warfare agent sarin. Using a comprehensive series of extended periodic models capable of capturing long-range sarin/water/framework interactions in both Zr6 and Zr12 MOFs, the effects of numbers and morphologies of defective sites, as well as ZrIV substitution with heavier CeIV, are thoroughly investigated. Our calculations show that hydrogen bonds involving both metal-oxide nodes and organic linkers can play important roles in the catalysis. Insights derived from this work should inform the design and realization of more effective and robust nextgeneration MOF-based heterogeneous catalysts.



INTRODUCTION Phosphorous-based chemical warfare agents (CWAs) are volatile chemicals known to inhibit the enzyme acetylcholine esterase (AChE) by phosphonylating its catalytic serine 203OH residue, thereby leading to the accumulation of the neurotransmitter acetylcholine in receptors, overstimulation, and ultimately death.1,2 First isolated from Pseudomonas diminuta, phosphotriesterase enzyme (PTE) with a positively charged bimetallic ZnII complex in its active site is shown to selectively hydrolyze different phosphotriesters including organophosphate pesticides and CWAs (Scheme 1a).3−5 Owing to the well-known high sensitivity of enzymes to environmental conditions and their fast deactivations, molecular engineering has been used to alter PTE to design alternative more robust (heterogeneous) biomimetic catalysts with high substrate specificity for practical real-world applications.6,7 The mechanistic details of PTE hydrolysis, for both wild-type and mutated enzymes, some including different transition metals in the active site, have been the subject of numerous experimental as well as theoretical studies.3,5,8−15 Nonenzymatic approaches to nerve agent hydrolysis are also of considerable interest. Metal−organic frameworks (MOFs) are promising porous organic/inorganic hybrid materials, which due to their robust chemical nature have found many applications in sorption, separation, and catalysis.16−20 Of particular interest for this work, missing linker-defective zirconium-based MOFs20−22 with adjacent open-metal sites © 2019 American Chemical Society

resembling those found in the active site of PTE have been shown to be highly active for the hydrolysis of CWAs in buffered aqueous media.7,23−37 Recently, we reported a detailed investigation of the hydrolysis mechanism of sarin, a CWA belonging to the G series of nerve agents that has been extensively studied experimentally, catalyzed by different Zrbased MOFs.38,39 With respect to those details, we found that for all studied Zr6-MOFs except MOF-808, the rate-determining step (RDS) for hydrolysis is the displacement of terminal Zr−OH2 groups by sarin, which explains why the experimentally observed hydrolysis rates increase upon a decrease in pH23 and/or dehydration24 of the Zr−SBUs (see Scheme 1b for a more complete mechanistic scheme).38 In another very recent study, we showed that a new [Zr6(μ3-O)4(μ3-OH)4(μ2-OH)6]2 double-node MOF has superior reactivity than various Zr6 single-node analogs. 39 The importance of adopting a reasonable computational model for studying the catalytic activity of these systems was also emphasized in our prior work, and we showed how reactivity trends for different MOFs change when different truncated clusters are employed as models.38,39 From a mechanistic point of view, the existence of openmetal sites, which often but not always originate from missing Received: April 23, 2019 Revised: May 21, 2019 Published: May 26, 2019 15157

DOI: 10.1021/acs.jpcc.9b03817 J. Phys. Chem. C 2019, 123, 15157−15165

Article

The Journal of Physical Chemistry C

Scheme 1. (a) Active Site of PTE,40 (b) Mechanistic Scheme for Hydrolysis of Sarin on ZrIV-MOFs, and (c) μ3-OH and μ3-O Faces of ZrIV Nodes Considered in Our Catalytic Study

Figure 1. Optimized structures for (a) MOF-808, (b) NU-1000 c pore, (c) benzene dicarboxylate linker, (d) benzene tricarboxylate linker, (e) pyrene tetracarboxylate linker, (f) benzene triphenyl dicarboxylate linker, (g) pristine UiO-66-12, (h) monodefective UiO-66-11, (i) bidefective UiO-66-10-I, (j) bidefective UiO-66-10-II, and (k) bidefective UiO-66-10-III. Generated empty pores after benzene dicarboxylate (BDC) linker removal are highlighted for different UiO-66 isomers. Legend: gray, white, red, and cyan represent C, H, O, and Zr atoms, respectively. See SI, Table S1, for relative energies of the different bidefective UiO-66 isomers, as well as optimized lattice constants of all studied MOFs.

linker defects, is crucial for most bond-breaking and bondforming reactions catalyzed by MOFs. Given the relative ease with which the extended framework may be structurally altered, either inadvertently or by design, under different experimental conditions such as temperature, pressure, solvent, and cosolvent or guest loadings, it is clearly important to take into account the presence of defects when designing realistic

materials for catalytic and/or other applications. The ubiquitous presence of defects in porous materials, especially MOFs, is fully embraced by the community, and indeed engineering them has recently become a topic of great interest to both theoretical and experimental groups, with the majority of studies focused on UiO-6641 as the archetype of stable Zr6O8 MOFs. Many interesting correlations between defects/ 15158

DOI: 10.1021/acs.jpcc.9b03817 J. Phys. Chem. C 2019, 123, 15157−15165

Article

The Journal of Physical Chemistry C structural disorder and observed chemical properties and activities have been reported.42−63 With respect to catalysis, which is the focus of the current study, understanding the intrinsic nature of individual defect sites can shed light on their propensity to be responsible for observed experimental activities. For example, Harvey et al. computed binding energies of different CWAs, including sarin, to Zr6-UiO-66 and Y6-UiO-66 MOFs focusing on the impact of different missing linker defect sites on computed energetics.64 While the adsorption of CWA is known to not be the RDS for the hydrolysis reaction,38,39 this contribution is the first study of its kind to shed light on the systematic nature and chemical properties of defect sites in robust MOFs as they interact with CWAs. It is known that the diffusion of substrates to the active site of different MOFs with different pore sizes and topologies in the presence of water molecules could potentially limit the rate of the reaction; however, this will be addressed in our future studies on the subject matter.65 Focusing on the complete hydrolysis reaction coordinate, in addition to identifying the correct proton topology of a newly synthesized Zr12 double-node MOF,66 we recently predicted higher reactivities for the hydrolysis of sarin by its bidefective framework than for any of its single-node analogues (Figure 2).38,39 Considering other variations in MOF composition, Farha et al.31 have shown that replacing ZrIV in UiO-66 with f-block CeIV transition metals can lead to enhanced reactivities for the hydrolysis of the CWA soman as well as the nerve agent simulant dimethyl 4-nitrophenyl phosphate (DMNP). Theoretical mechanistic studies have not yet been reported for the effects of metal substitution on the catalytic activity of this MOF or others. Our objective in this work is twofold: first, through performing systematic, dispersion-corrected quantum mechanical calculations on periodic models of different single- vs double-node ZrIV-MOFs (Figures 1 and 2), we aim to understand the impact of the number and morphology of defect sites on the reactivity of different MOFs in sarin hydrolysis as a test case. Having atomistic details for the effects of defect sites on reactivity in different MOFs may inform defect engineering opportunities to achieve higher reactivities or selectivities. Second, we aim to assess the effects of replacing ZrIV with CeIV f-block transition metals in catalytic hydrolysis of sarin in both UiO-66 and MOF-808. Results presented in this work can then serve as a benchmark for future theoretical studies, as well as guide experimental design and synthesis of more reactive MOF-based heterogeneous catalysts for specific applications.

Figure 2. Optimized crystal structures of (a) pristine Zr12, (b) monodefective Zr12, (c) bidefective Zr12, (d) tetradefective Zr12-I, (e) tetradefective Zr12-II, and (f) tetradefective Zr12-III double-node Zr12 MOFs. Generated empty pores after triphenyl benzene dicarboxylate (BDC) linker removal are highlighted. Legend: gray, white, red, and cyan represent C, H, O, and Zr atoms, respectively. See the Supporting Information (SI), Table S1, for relative energies of the different tetradefective Zr12 isomers, as well as optimized lattice constants of all studied MOFs.

electrons of the CeIV atoms.71 The plane-wave cutoff of the finest grid and REL_CUTOFF were set to 360 and 60 Ry, respectively. The default value of 10−5 Ry was used for all SCF convergences. Both atomic positions and cell parameters were relaxed for all pristine and defective single- and double-node MOFs but lattice parameters were kept fixed when locating intermediates and first-order saddle points, a valid assumption since the crystal structures of all studied MOFs are not expected to change much after loading the sarin substrate. All first-order saddle points along the reaction path of interest were located using the climbing image nudged elastic band (CI-NEB) method.72 All CI-NEB calculations consisted of 10 replicas for all studied single- and double-node MOFs. Changes in the electronic structure of all systems were examined by analyzing restrained electrostatic potential (RESP) charges calculated at the PBE-D3/DZVP-MOLOPT level, using the repeating electrostatic potential extracted atomic (REPEAT) method,73 especially benchmarked for porous crystalline solids such as MOFs. Details of the computations as well as Crystallography Information Files (CIFs) of all optimized crystal structures are included as part of the SI.



COMPUTATIONAL DETAILS All calculations were performed using periodic boundary condition at the PBE67 level with damped D3 dispersion correction68 using the hybrid Gaussian and plane-wave formalism, as implemented in the CP2K/quickstep package.69 The molecularly optimized double-ζ valence with polarization DZVP-MOLOPT basis sets and core−electron pseudopotentials according to the Geodecker−Teter−Hutter formulation70 were used in these calculations. The H (1s), C (2s, 2p), O (2s, 2p), F (2s, 2p), P (3s, 3p), and Zr (4s, 5s, 4p, 4d) electrons were treated as valence. For all CeIV-MOFs, spin-polarized calculations were performed using a modified pseudopotential and basis set in combination with the DFT+U method with the U value set to 7.0 eV to properly account for the localized 4f



RESULTS AND DISCUSSION Number and Morphology of Missing Linker Defects in Single- vs Double-Node ZrIV-MOFs and Their Impact on the Reactivity for Sarin Hydrolysis. Our previous work identified three elementary steps for the catalytic hydrolysis of sarin, namely, (i) water displacement by sarin, (ii) nucleophilic attack of water on the PO bond, and (iii) hydrogen fluoride (HF) elimination. Step (i) was computed to be the RDS for all 15159

DOI: 10.1021/acs.jpcc.9b03817 J. Phys. Chem. C 2019, 123, 15157−15165

Article

The Journal of Physical Chemistry C of the single- and double-node ZrIV-MOFs we examined except for MOF-808 (Scheme 1).38,39 However, we observed significant variations in computed energetics as a function of the size of capping groups in truncated cluster models, with the increasing size showing improved fidelity with fully periodic calculations, suggesting that the latter model is to be preferred when computationally tractable, and results discussed here are all from periodic calculations. Another important technical point is to note that, owing to the chirality of sarin and the different faces of the single- and double-node ZrIV metal-oxide nodes, many alternative local minima must be surveyed for every minimum-energy and transition-state (TS) structure (CIFs for all optimized structures are included as part of the SI). In the case of NU1000, we consider coordination only to “c pore” active sites since as shown in our previous study,38 the c pores are more active than the large pores in this MOF, presumably due to the existence of rather strong dispersion-like interactions in the narrower c pores than those in the large ones. This also agrees with recent studies reporting larger heats of adsorption for small substrates,74 as well as selective precursor deposition in atomic layer deposition75,76 in the c pores of this MOF. We further consider mono- and bidefective UiO-66, as well as mono-, bi-, and tetradefective Zr12 MOFs. Also, to examine the effects of changes in the morphology of defect sites on reactivity, three different isomers were considered for both bidefective UiO-66-10 and tetradefective Zr12 MOFs, with the isomers I−III spanning in relative electronic energies of 5.7 kcal/mol for the former (Figure 1) and 3.7 kcal/mol for the latter (Figure 2). The Computational Details section above provides additional methodological details. In concordance with our previous theoretical studies,38,39 we find here that H2O displacement by sarin is the RDS for all studied single- and double-node ZrIV-MOFs. We note that in our previous mechanistic study of MOF-808 using formateand benzoate-capped cluster models, we found otherwise, and predicted water nucleophilic attack to be rate-determining.38 We attribute this discrepancy to the deficiencies associated with using truncated cluster models, which neglect long-range framework−substrate interactions present in extended materials such as MOFs. Considering entirely the periodic results presented here, the computed energies of activation for monodefective UiO-66-11, bidefective UiO-66-10-I, bidefective UiO-66-10-II, bidefective UiO-66-10-III, NU-1000 c pore, and MOF-808 are 26.8, 28.1, 21.7, 27.1, 26.1, and 21.9 kcal/ mol, respectively (Figure 3 and SI Table S1). These data agree with the experiment that finds MOF-808 to be the most active of these catalysts for sarin hydrolysis, followed by NU-1000 and UiO-66, considering both monodefective UiO-66-11 and bidefective UiO-66-10-I.25 Meanwhile, the computed energies of activation associated with water displacement for mono-, bi-, and tetradefective Zr12 isomers I, II, and III are 20.0, 25.0, 35.0, 18.1, and 28.8 kcal/mol, respectively (Figure 4 and SI Table S1). The substantially greater variation in activation energy as a function of defect morphology in the Zr12 MOF compared to that in UiO-66 is interesting, as is the prediction that tetradefective Zr12-II should be more active than any Zr6 MOF given its low predicted activation energy of 18.1 kcal/ mol. To date, this MOF has not been tested for sarin hydrolysis. In the optimized displacement TS structures (SI Figure S1), the forming Zr−O(P) bond is the longest for MOF-808 and monodefective Zr12, i.e., most reactant-like, and shortest in

Figure 3. Periodic PBE-D3/DZVP-MOLOPT computed lowest energy paths for catalytic hydrolysis of sarin on different singlenode ZrIV-MOFs relative to their separated reactants. Key bond distances (in Å) are given for MOF-808 optimized displacement (TSdisplace) and addition (TSadd) transition states; all optimized structures are given in SI Figure S1. See Scheme 1b for the structures. Legend: gray, white, red, pink, dark cyan, and cyan represent C, H, O, P, F, and Zr atoms, respectively.

Figure 4. Periodic PBE-D3/DZVP-MOLOPT computed lowest energy paths for catalytic hydrolysis of sarin on different doublenode ZrIV-MOFs relative to their separated reactants. Key bond distances (in Å) are given for monodefective Zr12 optimized displacement (TSdisplace) and addition (TSadd) transition states. See Scheme 1b for the structures; all optimized structures are given in SI Figure S1. Legend: gray, white, red, pink, dark cyan, and cyan represent C, H, O, P, F, and Zr atoms, respectively.

bidefective UiO-66-10-III and tetradefective Zr12-I. This is consistent with the Bell−Evans−Polanyi principle77,78 that a faster reaction is associated with a TS structure that is more reactant-like. For most of the studied MOFs, two different, alternative first-order saddle points were located for the water displacement step in our climbing image nudged elastic band (CI-NEB) calculations, a lower-energy TS, which is structurally similar to the reactant, hereafter referred to as the early TS, 15160

DOI: 10.1021/acs.jpcc.9b03817 J. Phys. Chem. C 2019, 123, 15157−15165

Article

The Journal of Physical Chemistry C

Figure 5. CI-NEB optimized TS structures for water displacement (TSdisplace, top) and nucleophilic addition (TSadd, bottom) in the NU-1000 c pore. Key geometric data, as well as their electronic energies (in kcal/mol) relative to the HB isomer, are given. See Scheme 1 above and SI Figure S1 for the analogous structures of other MOFs. Legend: gray, white, red, yellow, green, and brown represent C, H, O, P, F, and Zr atoms, respectively.

(species “P” in Scheme 1b) by water and recovery of the original MOF active site, using our periodic models. PBE-D3 computed ΔEs for monodefective UiO-66-11 and bidefective UiO-66-10-I and MOF-808 are −2.3, −4.5, and −10.6 kcal/ mol, respectively, which shows that this step is exothermic and does contribute to the overall reactivity of the MOF but not to the turn-over frequency of these systems. Returning to the question of how the number and morphologies of the missing linker defect sites may alter the reactivity of these materials, we note the existence of an important trend in Figures 3 and 4. In both UiO-66 and the double-node Zr12 MOF, the most active species is the one characterized by multiple defects. Thus, for UiO-66, bidefective isomer II is maximally active, and it is isomer II of tetradefective Zr12-MOF that is also maximally active. However, the other isomers of these MOFs are predicted to be substantially less active than less heavily defected nodes, i.e., the monodefective nodes in each case. Considering the UiO-66 and double-node Zr12 cases in more depth, it is apparent that the relative orientations of the two defect sites play a decisive role in the activity. When the defect sites are on opposite faces of a single Zr6 component node (isomers III, in each case), high activation energies appear to be associated by a particularly stable H-bonded precomplex (cf. Figures 3 and 4) that increases the activation energy for water displacement with particularly short hydrogen bonds

and a second higher-energy TS resembling the product, which we call the late TS (following Hammond’s postulate;79 see Figure 5 above for NU-1000 c pore TSs and SI Figure S1 for all optimized TS structures). Interestingly, we located two alternative TSs for each water addition step as well: a lowenergy TS in which the attacking water molecule is located above the μ3-OH group of the metal-oxide node and accepts a hydrogen bond from it, and a second TS in which the attacking nucleophile is above the μ3-O group on the “other side” of the same node face and has no specific interaction with it (see Scheme 1c and Figure 5). Relative energies of these isomeric TSs were found to be 4.8 and 12.4 kcal/mol in favor of the hydrogen-bonded TS structures for NU-1000 c pore and MOF-808, respectively. The substantial stabilizing influence of the hydrogen bond is noteworthy as it would be expected to decrease the “absolute” nucleophilicity of the oxygen atom, but evidently, that effect is outweighed in the confined space of the reaction environment. Hydrogen bonding also discriminates the TS structures for water displacement in the NU-1000 c pore (Figure 5). In the early TS, the departing water forms two hydrogen bonds with a terminal OH and a pyrene carboxylate linker, whereas in the late TS, the water forms only one hydrogen bond with a terminal hydroxyl group. This leads to a 7.6 kcal/mol difference in energies for these two TS structures. We also computed the energetics for the last step of the catalytic cycle, i.e., replacing the HF-eliminated product 15161

DOI: 10.1021/acs.jpcc.9b03817 J. Phys. Chem. C 2019, 123, 15157−15165

Article

The Journal of Physical Chemistry C between water and terminal OH groups ranging from 1.442 Å in UiO-66-10-III to 1.415 Å in tetradefective Zr12-III (see SI Table S2). When the two sites share a single Zr atom around one meridian of a Zr6 core (isomers I, in each case), which too leads to very high water displacement activation energies, now primarily associated with rather weak hydrogen bonds between the O(P) of sarin and μ3-OH group of the metal-oxide node in the TS structures themselves, this hydrogen bond varies from 3.108 Å in isomer I to 2.158 Å in II and 2.619 Å in III in the case of the tetradefective Zr12 systems, agreeing perfectly with their computed order of activation energies (cf. Figures 3 and 4 and SI Table S2 again). However, when the two defects do not share a common Zr atom, as is true in both isomers II, there is a favorable destabilization of the initial HB complex and a stabilization of the TS structure that leads to especially low energy for water displacement. The correlation between hydrogen bond strength and reactivity observed in this work resonates well with a very recent study on a variety of structurally modified UiO-66 MOFs, in which the existence of a good correlation between the measured μ3-OH infrared frequencies of the modified UiO-66 nodes and electronic properties and reactivities in this MOF was reported.80 To further investigate the effects of stabilizing hydrogen bond interactions on the observed electronic structure and reactivity of different single- and double-node MOFs, we herein discuss our computed restrained electrostatic potential (RESP) charges on both precomplex and water displacement TS structures (see SI Table S3). In line with the observed hydrogen bond distances discussed above, the magnitude of the computed RESP charges on O(μ3‑OH), H(μ3‑OH) and terminal OH groups in isomers I−III of both UiO-66-10 and tetradefective Zr12 increases from isomers I−II and then decreases again in isomer III, all pointing toward the existence of stronger hydrogen bond interactions in isomers II of both the single-node UiO-66-10 and the double-node Zr12 MOFs, which agrees perfectly with their computed activation energies (cf. SI Tables S2 and S3). The takeaway from this analysis is that, while the construction of a material characterized by only monodefects might be expected to deliver a “good” catalyst, increasing the defecting level can give a better catalyst, but only if the “right” higher-order defects can be preferentially introduced, which certainly poses interesting synthetic challenges that we hope the experimental community will take up. Reactivity of CeIV vs ZrIV Single-Node MOFs in Sarin Hydrolysis. Owing to their high thermal, chemical, and mechanical robustness, CeIV analogues of Zr6-MOFs have begun to attract interest for a variety of applications.31,80,81 While some interest has derived from the potential to exploit the CeIII/CeIV couple for electro- or photochemical purposes, even in the absence of redox processes, one may expect the different Lewis acidity of CeIV compared to that of ZrIV together with different local structures to influence reactions catalyzed at the metal site. Indeed, Farha et al.31 have reported the catalytic activity of CeIV-BDC MOFs for the hydrolysis of CWAs and DMNP simulant. Here, we report the reactivity of the bidefective Ce-UiO-66-10-I and Ce-MOF-808 compared to that of their Zr analogues following the usual mechanism. Both Ce-substituted MOFs showed superior activities for hydrolysis of sarin compared to their Zr analogues (Figure 6). Specifically, upon Ce substitution, computed ΔE‡ values decreased from 28.1 to 14.3 kcal/mol for bidefective UiO66-10-I and from 21.9 to 11.0 kcal/mol for MOF-808.

Figure 6. Periodic PBE-D3/DZVP-MOLOPT computed lowest energy paths for catalytic hydrolysis of sarin (in kcal/mol) on different single-node ZrIV- vs CeIV bidefective UiO-66-10-I and MOF808 relative to their corresponding separated reactants. Key bond distances (in Å) are given for CeIV-MOF-808 optimized displacement (TSdisplace) and addition (TSadd) transition states; see Scheme 1b for the structures; all optimized structures are given in SI Figure S1. Legend: gray, white, red, pink, dark cyan, and yellow represent C, H, O, P, F, and Ce atoms, respectively.

Interestingly, most of the acceleration is associated with the facility with which water is displaced in the Ce case compared to that in Zr, reflecting the weaker bonding of water to the Ce node that has been documented experimentally.82,83 Our computed RESP charges of sarin in displacement TSs show that the PO bond becomes more polarized upon adsorption on CeIV-MOF-808 than on its ZrIV analogue with Δq on O and P being −0.029 and +0.042 e (see SI Table S3). Interestingly, the oxygen atom of the attacking water molecule also carries a more negative charge (Δq = −0.053 e) when it is on CeIVMOF-808 node than on its ZrIV analogue, which signals the superior nucleophilicity of water molecules on the CeIV nodes than that on the ZrIV ones (SI Table S3). Overall, our calculations show CeIV-MOF-808 to be the fastest of all studied MOFs for the hydrolysis of sarin. We note that the functionalization of the linkers of CeIV-MOFs with, for example, −NH2 functional groups (that have been shown to enhance reactivity for their ZrIV analogues7,30,32,34,36) and recently iodine groups84 could offer additional opportunity to further enhance the reactivity of these systems. Such predictions await experimental verification.



CONCLUSIONS In summary, using fully periodic models and PBE functional with D3 dispersion correction, we investigated in detail the reactivity of different single- and double-node MOFs in the hydrolysis of the sarin nerve agent as a test case. Through careful investigation of different factors including the number and morphology of the missing linker defect sites, we found that more defects can lead to lower or higher activity, depending on how they are arranged relative to one another. This poses an interesting synthetic challenge to the experimental community for making highly active MOFbased heterogeneous catalysts by engineering defects in them. ZrIV substitution with CeIV was found to reduce the computed activation energies for bidefective UiO-66-10-I and MOF-808 15162

DOI: 10.1021/acs.jpcc.9b03817 J. Phys. Chem. C 2019, 123, 15157−15165

Article

The Journal of Physical Chemistry C

(5) Caldwell, S. R.; Raushel, F. M.; et al. Primary and Secondary Oxygen-18 Isotope Effects in the Alkaline and Enzyme-Catalyzed Hydrolysis of Phosphotriesters. J. Am. Chem. Soc. 1991, 113, 730− 132. (6) Hill, C. M.; Li, W.-S.; Thoden, J. B.; Holden, H. M.; Raushel, F. M. Enhanced Degradation of Chemical Warfare Agents through Molecular Engineering of the Phosphotriesterase Active Site. J. Am. Chem. Soc. 2003, 125, 8990−8991. (7) Katz, M. J.; Mondloch, J. E.; Totten, R. K.; Park, J. K.; Nguyen, S. T.; Farha, O. K.; Hupp, J. T. Simple and Compelling Biomimetic Metal−Organic Framework Catalyst for the Degradation of Nerve Agent Simulants. Angew. Chem., Int. Ed. 2014, 53, 497−501. (8) Koča, J.; Zhan, C.-G.; Rittenhouse, R. C.; Ornstein, R. L. Mobility of the Active Site Bound Paraoxon and Sarin in ZincPhosphotriesterase by Molecular Dynamics Simulation and Quantum Chemical Calculation. J. Am. Chem. Soc. 2001, 123, 817−826. (9) Krauss, M.; Olsen, L.; Antony, J.; Hemmingsen, L. Coordination Geometries of Zn(II) and Cd(II) in Phosphotriesterase: Influence of Water Molecules in the Active Site. J. Phys. Chem. B 2002, 106, 9446− 9453. (10) Aubert, S. D.; Li, Y.; Raushel, F. M. Mechanism for the Hydrolysis of Organophosphates by the Bacterial Phosphotriesterase. Biochemistry 2004, 43, 5707−5715. (11) Jackson, C.; Kim, H.-K.; Carr, P. D.; Liu, J.-W.; Ollis, D. L. The structure of an enzyme−product complex reveals the critical role of a terminal hydroxide nucleophile in the bacterial phosphotriesterase mechanism. Biochim. Biophys. Acta, Proteins Proteomics 2005, 1752, 56−64. (12) Chen, S.-L.; Fang, W.-H.; Himo, F. Theoretical Study of the Phosphotriesterase Reaction Mechanism. J. Phys. Chem. B 2007, 111, 1253−1255. (13) Chen, S.-L.; Fang, W.-H.; Himo, F. Technical aspects of quantum chemical modeling of enzymatic reactions: the case of phosphotriesterase. Theor. Chem. Acc. 2008, 120, 515−522. (14) Perezgasga, L.; Sánchez-Sánchez, L.; Aguila, S.; VazquezDuhalt, R. Substitution of the Catalytic Metal and Protein PEGylation Enhances Activity and Stability of Bacterial Phosphotriesterase. Appl. Biochem. Biotechnol. 2012, 166, 1236−1247. (15) Bigley, A. N.; Raushel, F. M. Catalytic Mechanisms for Phosphotriesterases. Biochim. Biophys. Acta, Proteins Proteomics 2013, 1834, 443−453. (16) Corma, A.; García, H.; Llabrés i Xamena, F. X. Engineering Metal Organic Frameworks for Heterogeneous Catalysis. Chem. Rev. 2010, 110, 4606−4655. (17) Farha, O. K.; Hupp, J. T. Rational Design, Synthesis, Purification, and Activation of Metal−Organic Framework Materials. Acc. Chem. Res. 2010, 43, 1166−1175. (18) Yoon, M.; Srirambalaji, R.; Kim, K. Homochiral Metal− Organic Frameworks for Asymmetric Heterogeneous Catalysis. Chem. Rev. 2012, 112, 1196−1231. (19) Cui, Y.; Li, B.; He, H.; Zhou, W.; Chen, B.; Qian, G. Metal− Organic Frameworks as Platforms for Functional Materials. Acc. Chem. Res. 2016, 49, 483−493. (20) Odoh, S.; Cramer, C. J.; Truhlar, D.; Gagliardi, L. QuantumChemical Characterization of the Properties and Reactivities of Metal−Organic Frameworks. Chem. Rev. 2015, 115, 6051−6111. (21) Furukawa, H.; Cordova, K. E.; O'Keeffe, O.; Yaghi, O. M. The Chemistry and Applications of Metal−Organic Frameworks. Science 2013, 341, No. 1230444. (22) Bai, Y.; Dou, Y.; Xie, L.-H.; Rutledge, W.; Li, J.-R.; Zhou, H.-C. Zr-Based Metal−Organic Frameworks: Design, Synthesis, Structure, and Applications. Chem. Soc. Rev. 2016, 45, 2327−2367. (23) Katz, M. J.; Klet, R. C.; Moon, S.-Y.; Mondloch, J. E.; Hupp, J. T.; Farha, O. K. One Step Backward Is Two Steps Forward: Enhancing the Hydrolysis Rate of UiO-66 by Decreasing [OH−]. ACS Catal. 2015, 5, 4637−4642. (24) Mondloch, J. E.; Katz, M. J.; Isley, W. C., III; Ghosh, P.; Liao, P.; Bury, W.; Wagner, G. W.; Hall, M. G.; DeCoste, J. B.; Peterson, G. W.; Snurr, R. Q.; Cramer, C. J.; Hupp, J. T.; Farha, O. K. Destruction

by half, which agrees qualitatively with the available experimental data on CeIV-BDC. Hydrogen bond formations with both the metal-oxide nodes and organic linkers were shown to play an important role in the decomposition of the sarin CWA. This study showcases the strength and importance of considering both defect engineering and metal substitution as a means for tuning the reactivity of MOF-based porous materials and presents opportunities for applications that make use of defect engineering.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.jpcc.9b03817. PBE-D3/DZVP-MOLOPT computed periodic lattice constants (Table S1); PBE-D3/DZVP-MOLOPT computed key geometric data (Table S2); PBE-D3/DZVPMOLOPT computed RESP charges (Table S3); and PBE-D3/DZVP-MOLOPT CI-NEB optimized displacement and addition TSs of different single-node ZrIV and CeIV MOFs (Figure S1) (PDF) (ZIP)



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Mohammad R. Momeni: 0000-0002-7731-5823 Christopher J. Cramer: 0000-0001-5048-1859 Present Address †

Department of Chemistry and Biochemistry, University of California, San Diego, La Jolla, California 92093, United States, [email protected] (M.R.M.). Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS The authors gratefully acknowledge the Defense Threat Reduction Agency (HDTRA1-18-1-0003) for financial support. The authors also acknowledge the Minnesota Supercomputing Institute (MSI) for providing resources that contributed to the research results reported within this paper. M.R.M. is grateful for helpful discussions with Omar Farha, Timur Islamoglu, Farnaz Shakib, and Manuel Ortuño.



REFERENCES

(1) Wilson, I. B.; Ginsburg, B. A powerful Reactivator of Alkylphosphate-Inhibited Acetylcholinesterase. Biochim. Biophys. Acta 1955, 18, 168−170. (2) Allgardsson, A.; Berg, L.; Akfur, C.; Hörnberg, A.; Worek, F.; Linusson, A.; Ekström, F. J. Structure of a Prereaction Complex Between the Nerve Agent Sarin, Its Biological Target Acetylcholinesterase, and the Antidote HI-6. Proc. Natl. Acad. Sci. U. S. A. 2016, 113, 5514−5519. (3) Lewis, V. E.; Donarski, W. J.; Wild, J. R.; Raushel, F. M. Mechanism and Stereochemical Course at Phosphorus of the Reaction Catalyzed by a Bacterial Phosphotriesterase. Biochemistry 1988, 27, 1591−1597. (4) Donarski, W. J.; Dumas, D. P.; Heitmeyer, D. P.; Lewis, V. E.; Raushel, F. M. Structure-Activity Relationships in the Hydrolysis of Substrates by the Phosphotriesterase from Pseudomonas Diminuta. Biochemistry 1989, 28, 4650−4655. 15163

DOI: 10.1021/acs.jpcc.9b03817 J. Phys. Chem. C 2019, 123, 15157−15165

Article

The Journal of Physical Chemistry C of Chemical Warfare Agents Using Metal−Organic Frameworks. Nat. Mater. 2015, 14, 512−516. (25) Moon, S.-Y.; Liu, Y.; Hupp, J. T.; Farha, O. K. Instantaneous Hydrolysis of Nerve-Agent Simulants with a Six-Connected Zirconium-Based Metal−Organic Framework. Angew. Chem., Int. Ed. 2015, 54, 6795−6799. (26) Plonka, A. M.; Wang, Q.; Gordon, W. O.; Balboa, A.; Troya, D.; Guo, W.; Sharp, C. H.; Senanayake, S. D.; Morris, J. R.; Hill, C. L.; Frenkel, A. I. In Situ Probes of Capture and Decomposition of Chemical Warfare Agent Simulants by Zr-Based Metal Organic Frameworks. J. Am. Chem. Soc. 2017, 139, 599−602. (27) Peterson, G. W.; Moon, S.-Y.; Wagner, G. W.; Hall, M. G.; DeCoste, J. B.; Hupp, J. T.; Farha, O. K. Tailoring the Pore Size and Functionality of UiO-Type Metal−Organic Frameworks for Optimal Nerve Agent Destruction. Inorg. Chem. 2015, 54, 9684−9686. (28) Katz, M. J.; Moon, S.-Y.; Mondloch, J. E.; Beyzavi, M. H.; Stephenson, C. J.; Hupp, J. T.; Farha, O. K. Exploiting Parameter Space in MOFs: A 20-Fold Enhancement of Phosphate-Ester Hydrolysis with UiO-66-NH2. Chem. Sci. 2015, 6, 2286−2291. (29) Liu, Y.; Moon, S.-Y.; Hupp, J. T.; Farha, O. K. Dual-Function Metal−Organic Framework as a Versatile Catalyst for Detoxifying Chemical Warfare Agent Simulants. ACS Nano 2015, 9, 12358− 12364. (30) Moon, S. Y.; Wagner, G. W.; Mondloch, J. E.; Peterson, G. W.; DeCoste, J. B.; Hupp, J. T.; Farha, O. K. Effective, Facile, and Selective Hydrolysis of the Chemical Warfare Agent VX Using Zr6Based Metal−Organic Frameworks. Inorg. Chem. 2015, 54, 10829− 10833. (31) Islamoglu, T.; Atilgan, A.; Moon, S.-Y.; Peterson, G. W.; DeCoste, J. B.; Hall, M.; Hupp, J. T.; Farha, O. K. Cerium(IV) vs Zirconium(IV) Based Metal−Organic Frameworks for Detoxification of a Nerve Agent. Chem. Mater. 2017, 29, 2672−2675. (32) DeCoste, J. B.; Peterson, G. W. Metal−Organic Frameworks for Air Purification of Toxic Chemicals. Chem. Rev. 2014, 114, 5695− 5727. (33) de Koning, M. C.; Van Grol, M.; Breijaert, T. Degradation of Paraoxon and the Chemical Warfare Agents VX, Tabun, and Soman by the Metal−Organic Frameworks UiO-66-NH2, MOF-808, NU1000, and PCN-777. Inorg. Chem. 2017, 56, 11804−11809. (34) Wang, G.; Sharp, C.; Plonka, A. M.; Wang, Q.; Frenkel, A. I.; Guo, W.; Hill, C.; Smith, C.; Kollar, J.; Troya, D.; Morris, J. R. Mechanism and Kinetics for Reaction of the Chemical Warfare Agent Simulant, DMMP(g), with Zirconium(IV) MOFs: An UltrahighVacuum and DFT Study. J. Phys. Chem. C 2017, 121, 11261−11272. (35) Gil-San-Millan, R.; López-Maya, E.; Hall, M.; Padial, N. M.; Peterson, G. W.; DeCoste, J. B.; Rodríguez-Albelo, L. M.; Oltra, J. E.; Barea, E.; Navarro, J. A. R. Chemical Warfare Agents Detoxification Properties of Zirconium Metal−Organic Frameworks by Synergistic Incorporation of Nucleophilic and Basic Sites. ACS Appl. Mater. Interfaces 2017, 9, 23967−23973. (36) Islamoglu, T.; Ortuño, M. A.; Proussaloglou, E.; Howarth, A. J.; Vermeulen, N. A.; Atilgan, A.; Asiri, A. M.; Cramer, C. J.; Farha, O. K. Presence versus Proximity: The Role of Pendant Amines in the Catalytic Hydrolysis of a Nerve Agent Simulant. Angew. Chem. 2018, 130, 1967−1971. (37) Troya, D. Reaction Mechanism of Nerve-Agent Decomposition with Zr-Based Metal Organic Frameworks. J. Phys. Chem. C 2016, 120, 29312−29323. (38) Momeni, M. R.; Cramer, C. J. Dual Role of Water in Heterogeneous Catalytic Hydrolysis of Sarin by Zirconium-Based Metal−Organic Frameworks. ACS Appl. Mater. Interfaces 2018, 10, 18435−18439. (39) Momeni, M. R.; Cramer, C. J. Structural Characterization of Pristine and Defective [Zr12(μ3-O)8(μ3-OH)8(μ2-OH)6]18+ DoubleNode Metal−Organic Framework and Predicted Applications for Single-Site Catalytic Hydrolysis of Sarin. Chem. Mater. 2018, 30, 4432−4439. (40) Graphics were generated with PyMol V. 2.0. The PyMOL Molecular Graphics System, version 2.0; Schrödinger, LLC, 2018.

(41) Cavka, J. H.; Jakobsen, S.; Olsbye, U.; Guillou, N.; Lamberti, C.; Bordiga, S.; Lillerud, C. P. A New Zirconium Inorganic Building Brick Forming Metal Organic Frameworks with Exceptional Stability. J. Am. Chem. Soc. 2008, 130, 13850−13851. (42) Bennett, T. D.; Cheetham, A. K.; Fuchs, A. H.; Coudert, F. X. Interplay Between Defects, Disorder and Flexibility in Metal-Organic Frameworks. Nat. Chem. 2017, 9, 11−16. (43) Fang, Z.; Bueken, B.; De Vos, D. E.; Fischer, R. A. DefectEngineered Metal−Organic Frameworks. Angew. Chem., Int. Ed. 2015, 54, 7234−7254. (44) Sholl, D. S.; Lively, R. P. Defects in Metal−Organic Frameworks: Challenge or Opportunity? J. Phys. Chem. Lett. 2015, 6, 3437−3444. (45) Cheetham, A. K.; Bennett, T. D.; Coudert, F.-X.; Goodwin, A. L. Defects and Disorder in Metal Organic Frameworks. Dalton Trans. 2016, 45, 4113−4126. (46) Dissegna, S.; Epp, K.; Heinz, W. R.; Kieslich, G.; Fischer, R. A. Defective Metal−Organic Frameworks. Adv. Mater. 2018, 30, No. 1704501. (47) Cliffe, M. J.; Wan, W.; Zou, X.; Chater, P. A.; Kleppe, A. K.; Tucker, M. G.; Wilhelm, H.; Funnell, N. P.; Coudert, F.-X.; Goodwin, A. L. Correlated Defect Nanoregions in a Metal−Organic Framework. Nat. Commun. 2014, 5, No. 4176. (48) Valenzano, L.; Civalleri, B.; Chavan, S.; Bordiga, S.; Nilsen, M. H.; Jakobsen, S.; Lillerud, K. P.; Lamberti, C. Disclosing the Complex Structure of UiO-66 Metal Organic Framework: A Synergic Combination of Experiment and Theory. Chem. Mater. 2011, 23, 1700−1718. (49) Cliffe, M. J.; Hill, J. A.; Murray, C. A.; Coudert, F.-X.; Goodwin, A. L. Defect-Dependent Colossal Negative Thermal Expansion in UiO-66(Hf) Metal−Organic Framework. Phys. Chem. Chem. Phys. 2015, 17, 11586−11592. (50) Thornton, A. W.; Babarao, R.; Jain, A.; Trousselet, F.; Coudert, F.-X. Defects in Metal−Organic Frameworks: A Compromise Between Adsorption and Stability? Dalton Trans. 2016, 45, 4352− 4359. (51) Zhang, C.; Han, C.; Sholl, D. S.; Schmidt, J. R. Computational Characterization of Defects in Metal−Organic Frameworks: Spontaneous and Water-Induced Point Defects in ZIF-8. J. Phys. Chem. Lett. 2016, 7, 459−464. (52) Bennett, T. D.; Todorova, T. K.; Baxter, E. F.; Reid, D. G.; Gervais, C.; Bueken, B.; Van de Voorde, B.; De Vos, D.; Keen, D. A.; Mellot-Draznieks, C. Connecting Defects and Amorphization in UiO66 and MIL-140 Metal−Organic Frameworks: A Combined Experimental and Computational Study. Phys. Chem. Chem. Phys. 2016, 18, 2192−2201. (53) Bueken, B.; Van Velthoven, N.; Krajnc, A.; Smolders, A.; Taulelle, F.; Mellot-Draznieks, C.; Mali, G.; Bennett, T. D.; De Vos, D. Tackling the Defect Conundrum in UiO-66: A Mixed-Linker Approach to Engineering Missing Linker Defects. Chem. Mater. 2017, 29, 10478−10486. (54) Yuan, L.; Tian, M.; Lan, J.; Cao, X.; Wang, X.; Chai, Z.; Gibson, J. K.; Shi, W. Defect Engineering in Metal−Organic Frameworks: A New Strategy To Develop Applicable Actinide Sorbents. Chem. Commun. 2018, 54, 370−373. (55) Liang, W.; Li, L.; Hou, J.; Shepherd, N. D.; Bennett, T. D.; D’Alessandro, D. M.; Chen, V. Linking Defects, Hierarchical Porosity Generation and Desalination Performance in Metal−Organic Frameworks. Chem. Sci. 2018, 9, 3508−3516. (56) Firth, F.; Cliffe, M.; Vulpe, D.; Moghadam, P.; Fairen-Jimenez, D.; Slater, B.; et al. Engineering New Defective Phases of UiO Family Metal−Organic Frameworks with Water. J. Mater. Chem. A 2019, 7, 7459−7469. (57) Park, H.; Kim, S.; Jung, B.; Park, M. H.; Kim, Y.; Kim, M. Defect Engineering into Metal−Organic Frameworks for the Rapid and Sequential Installation of Functionalities. Inorg. Chem. 2018, 57, 1040−1047. 15164

DOI: 10.1021/acs.jpcc.9b03817 J. Phys. Chem. C 2019, 123, 15157−15165

Article

The Journal of Physical Chemistry C (58) Fu, Y.; Kang, Z.; Yin, J.; Cao, W.; Tu, Y.; Wang, Q.; Kong, X. Duet of Acetate and Water at the Defects of Metal−Organic Frameworks. Nano Lett. 2019, 19, 1618−1624. (59) Vandichel, M.; Hajek, J.; Vermoortele, F.; Waroquier, M.; De Vos, D. E.; Van Speybroeck, V. Active Site Engineering in UiO-66 Type Metal−Organic Frameworks By Intentional Creation of Defects: A Theoretical Rationalization. CrystEngComm 2015, 17, 395−406. (60) Rogge, S. M. J.; Wieme, J.; Vanduyfhuys, L.; Vandenbrande, S.; Maurin, G.; Verstraelen, T.; Waroquier, M.; Van Speybroeck, V. Thermodynamic Insight in the High-Pressure Behavior of UiO-66: Effect of Linker Defects and Linker Expansion. Chem. Mater. 2016, 28, 5721−5732. (61) Vandichel, M.; Hajek, J.; Ghysels, A.; De Vos, A.; Waroquier, M.; Van Speybroeck, V. Water Coordination and Dehydration Processes in Defective UiO-66 Type Metal Organic Frameworks. CrystEngComm 2016, 18, 7056−7069. (62) Marreiros, J.; Caratelli, C.; Hajek, J.; Krajnc, A.; Fleury, G.; Bueken, B.; De Vos, D. E.; Mali, G.; Roeffaers, M. B. J.; Van Speybroeck, V.; Ameloot, R. Active Role of Methanol in PostSynthetic Linker Exchange in the Metal−Organic Framework UiO-66. Chem. Mater. 2019, 31, 1359−1369. (63) Svane, K. L.; Bristow, J. K.; Gale, J. D.; Walsh, A. Vacancy Defect Configurations in the Metal−Organic Framework UiO-66: Energetics and Electronic Structure. J. Mater. Chem. A 2018, 6, 8507− 8513. (64) Harvey, J. A.; Greathouse, J. A.; Sava-Gallis, D. F. Defect and Linker Effects on the Binding of Organophosphorous Compounds in UiO-66 and Rare-Earth MOFs. J. Phys. Chem. C 2018, 122, 26889− 26896. (65) Skoulidas, A. I.; Sholl, D. S. Self-Diffusion and Transport Diffusion of Light Gases in Metal-Organic Framework Materials Assessed Using Molecular Dynamics Simulations. J. Phys. Chem. B 2005, 109, 15760−15768. (66) Ji, P.; Manna, K.; Lin, Z.; Feng, X.; Urban, A.; Song, Y.; Lin, W. Single-Site Cobalt Catalysts at New Zr12(μ3-O)8(μ3-OH)8(μ2-OH)6 Metal−Organic Framework Nodes for Highly Active Hydrogenation of Nitroarenes, Nitriles, and Isocyanides. J. Am. Chem. Soc. 2017, 139, 7004−7011. (67) Perdew, J.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1997, 78, No. 1396. (68) Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A Consistent and Accurate ab Initio Parametrization of Density Functional Dispersion Correction (DFT-D) for the 94 Elements H-Pu. J. Chem. Phys. 2010, 132, No. 154104. (69) Hutter, J.; Iannuzzi, M.; Schiffmann, F.; VandeVondele, J. CP2K: Atomistic Simulations of Condensed Matter Systems. Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2014, 4, 15−25. (70) Goedecker, S.; Teter, M.; Hutter, J. Separable Dual-Space Gaussian Pseudopotentials. Phys. Rev. B 1996, 54, 1703−1710. (71) Wang, Y.-G.; Mei, D.; Li, J.; Rousseau, R. DFT+U Study on the Localized Electronic States and Their Potential Role During H2O Dissociation and CO Oxidation Processes on CeO2(111) Surface. J. Phys. Chem. C 2013, 117, 23082−23089. (72) Henkelman, G.; Uberuaga, B. P.; Jónsson, H. A Climbing Image Nudged Elastic Band Method For Finding Saddle Points and Minimum Energy Paths. J. Chem. Phys. 2000, 113, 9901−9904. (73) Campaña,́ C.; Mussard, B.; Woo, T. K. Electrostatic Potential Derived Atomic Charges for Periodic Systems Using a Modified Error Functional. J. Chem. Theory Comput. 2009, 5, 2866−2878. (74) Zhang, W.; Ma, Y.; Santos-Loṕez, I. A.; Lownsbury, J. M.; Yu, H.; Liu, W.-G.; Truhlar, D. G.; Campbell, C. T.; Vilches, O. E. Energetics of van der Waals Adsorption on the Metal−Organic Framework NU-1000 with Zr6-oxo, Hydroxo, and Aqua Nodes. J. Am. Chem. Soc. 2018, 140, 328−338. (75) Gallington, L. C.; Kim, I. S.; Liu, W.-G.; Yakovenko, A. A.; Platero-Prats, A. E.; Li, Z.; Wang, T. C.; Hupp, J. T.; Farha, O. K.; Truhlar, D. G.; Martinson, A. B. F.; Chapman, K. W. Regioselective

Atomic Layer Deposition in Metal−Organic Frameworks Directed by Dispersion Interactions. J. Am. Chem. Soc. 2016, 138, 13513−13516. (76) Liu, W.-G.; Truhlar, D. G. Computational Linker Design for Highly Crystalline Metal−Organic Framework NU-1000. Chem. Mater. 2017, 29, 8073−8081. (77) Bell, R. P. The Theory of Reactions Involving Proton Transfers. Proc. R. Soc. London, Ser. A 1936, 154, 414−429. (78) Evans, M. G.; Polanyi, M. Further Considerations on the Thermodynamics of Chemical Equilibria and Reaction Rates. Trans. Faraday Soc. 1936, 32, 1333−1360. (79) Hammond, G. S. A. Correlation of Reaction Rates. J. Am. Chem. Soc. 1955, 77, 334−338. (80) Wei, R.; Gaggioli, C. A.; Li, G.; Islamoglu, T.; Zhang, Z.; Yu, P.; Farha, O. K.; Cramer, C. J.; Gagliardi, L.; Yang, D.; Gates, B. C. Tuning the Properties of Zr6O8 Nodes in the Metal Organic Framework UiO-66 By Selection of Node-Bound Ligands and Linkers. Chem. Mater. 2019, 31, 1655−1663. (81) Wu, X.-P.; Gagliardi, L.; Truhlar, D. G. Cerium Metal−Organic Framework for Photocatalysis. J. Am. Chem. Soc. 2018, 140, 7904− 7912. (82) Lammert, M.; Wharmby, M. T.; Smolders, S.; Bueken, B.; Lieb, A.; Lomachenko, K. A.; De Vosc, K.; Stock, N. Cerium-Based Metal Organic Frameworks with UiO-66 Architecture: Synthesis, Properties and Redox Catalytic Activity. Chem. Commun. 2015, 51, 12578− 12581. (83) Islamoglu, T.; Ray, D.; Li, P.; Majewski, M. B.; Akpinar, I.; Zhang, Z.; Cramer, C. J.; Gagliardi, L.; Farha, O. K. From Transition Metals to Lanthanides to Actinides: Metal-Mediated Tuning of Electronic Properties of Isostructural Metal−Organic Frameworks. Inorg. Chem. 2018, 57, 13246−13251. (84) Kalaj, M.; Momeni, M. R.; Bentz, K. C.; Barcus, K. S.; Palomba, J. M.; Paesani, F.; Cohen, S. M. Halogen Bonding in UiO-66 Frameworks Promotes Superior Chemical Warfare Agent Simulant Degradation. Chem. Commun. 2019, 55, 3481−3484.

15165

DOI: 10.1021/acs.jpcc.9b03817 J. Phys. Chem. C 2019, 123, 15157−15165