Stability of Zeolitic Imidazolate Frameworks in NO2 - The Journal of

Jan 4, 2019 - How Useful Are Common Simulants of Chemical Warfare Agents at Predicting Adsorption Behavior? The Journal of Physical Chemistry C...
1 downloads 0 Views 1MB Size
Subscriber access provided by UNIV OF LOUISIANA

C: Surfaces, Interfaces, Porous Materials, and Catalysis 2

Stability of Zeolitic Imidazolate Frameworks in NO Souryadeep Bhattacharyya, Rebecca Han, Jayraj N. Joshi, Guanghui Zhu, Ryan P Lively, Krista S. Walton, David S. Sholl, and Sankar Nair

J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.8b11377 • Publication Date (Web): 04 Jan 2019 Downloaded from http://pubs.acs.org on January 12, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Stability of Zeolitic Imidazolate Frameworks in NO2 Souryadeep Bhattacharyya, Rebecca Han, Jayraj N. Joshi, Guanghui Zhu, Ryan P. Lively, Krista S. Walton, David S. Sholl*, and Sankar Nair*

School of Chemical & Biomolecular Engineering, Georgia Institute of Technology 311 Ferst Drive NW, Atlanta, GA 30318, USA

* Corresponding authors: [email protected], [email protected]

1 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 36

Abstract The use of nanoporous zeolitic imidazolate frameworks (ZIFs) in separation processes is attractive, but the presence of acid gases such as SOx or NOx in process streams can have detrimental effects. While we have recently developed a mechanistic picture of SO x-ZIF interactions, here we describe the remarkably different effects of NO2 on ZIFs. ZIFs with a representative range of SOx stabilities are all unstable - as defined by loss of crystallinity and porosity - in dry and humid NO2, whereas most ZIFs are stable in dry SOx and some even in humid SOx. A detailed mechanism is developed based upon FTIR spectroscopy and DFT calculations. H-abstraction by free radical NO2 and subsequent acidic species formation are the major degradation pathways, while adsorbed HNO3 formation in humid conditions is an additional pathway in hydrophilic ZIFs. These findings strongly suggest that new strategies to stabilize ZIFs/MOFs towards NO2 attack are required.

2 ACS Paragon Plus Environment

Page 3 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Introduction Nitrogen oxides (NOx) are harmful air pollutants whose toxic effects are attributed primarily to NO2, which also enables formation of ozone and particulate matter.1-3 NOx removal from point sources such as automobiles or power plants is often carried out by selective catalytic reduction3-5 which is energy-intensive, requires large volumes of catalyst and reductant (usually NH3), and is susceptible to catalyst deactivation. A potential alternative is by NOx adsorption in porous materials. In addition to conventional activated carbons, metal oxides, and zeolites, microporous metal-organic frameworks (MOFs) such as zeolitic imidazolate frameworks (ZIFs) could be used.3 ZIFs consist of metal centers (Zn2+/Co2+) tetrahedrally coordinated to imidazolate linkers.6-7 Judicious selection of linkers and synthesis conditions allows tuning of their pore apertures and cages.7-8 ZIF stability in acid gases such as NO2 is a critical consideration for their use in environments containing these gases.9-14 Many ZIFs (and other MOFs) have been shown to have high kinetic stability in dry SOx, although fewer are stable in humid SOx.10-12 However, much less is known about interactions of ZIFs/MOFs with NOx acid gases.13-14

The mechanisms of NO2 interactions with MOFs can be quite different from SOx, so that previous findings on SOx stability may not be applicable. The open-metal site MOF CuBTC was reported to be unstable in dry or humid NO2.1,

15

Under dry conditions, formation of nitrates

bound to Cu and release of NO was proposed as the reaction mechanism. The coordinatively saturated MOF UiO-6616-19 retained its surface area under humid NOx (71% RH and 1000 ppm NO2) but degraded in dry NO2, while the more hydrophilic UiO-67 degraded under both conditions. Water was proposed to competitively prevent NO2 adsorption in UiO-66, but was proposed to dissolve NO2 and form nitric acid that attacked the hydrophilic UiO-67. ZIF-7, ZIF-

3 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 36

8 and ZIF-90 were exposed to 15 ppm SO2 followed by 10 ppm NO2 at 80% RH for 2 days each and then characterized,20 and so the effects of the individual gases could not be distinguished. To our knowledge there are no other reports on the effects of NO2 on ZIF materials.

Here we show that the mechanisms of ZIF interaction with (and degradation by) NO2 are strikingly different from the mechanisms of their interaction with CO2 or SOx. We systematically investigate the reaction mechanisms of dry and humid NO2 with ZIFs and their impact on the stability of these materials. The material stability is defined in terms of the retention of its crystal structure and pore volume.21 The main conclusions of our experimental investigation are also supported by a detailed computational study of defect formation energies, highlighting the role of computational modeling in understanding the acid gas stability of ZIFs.22-25 Many of the mechanistic findings of this work are expected to be applicable to a wider variety of MOFs. Our recent findings on SOx stability of a large set of more than 15 ZIFs10 allow us to choose three representative ZIFs for the present work, that encompass different outcomes possible upon NO2 exposure. ZIF-8 and ZIF-90 were selected because they had the highest and lowest measurable degradation rates in humid SOx. ZIF-71 was selected because it showed no measurable degradation in humid SOx. The topological characteristics of the three ZIFs are shown in Table S1 (Supporting Information).

Materials and Methods Materials Zinc(II) nitrate hexahydrate (99 %) and imidazole-2-carboxaldehyde (97 %) were obtained from Alfa Aesar. 2-methylimidazole (99 %), 4,5-dichloroimidazole (98 %) and sodium 4 ACS Paragon Plus Environment

Page 5 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

formate (99 %) were purchased from Sigma-Aldrich while methanol (99.8 %) and N,Ndimethylformamide (DMF) (99.8 %) were purchased from BDH. Chemicals were used as received. Deionized (DI) water from the EMD Millipore water purification system and ultra-high purity air (76.5-80.5 % N2, 19.5-23.5 % O2) from Airgas were used in this work.

ZIF Synthesis The synthesis protocol reported by Gee et al was modified to synthesize ZIF-8 and ZIF90 in this work.

26

For ZIF-8 synthesis, a solution of 0.972 g 2-methylimidazole and 1.614 g

sodium formate in 120 cc methanol was mixed with another solution of 1.764 g zinc (II) nitrate in 120 cc methanol, followed by heating at 363 K for 24 hours. The collected crystals were washed three times with methanol and air dried at 333 K. ZIF-90 synthesis was carried out by adding 11.904 g zinc (II) nitrate and 15.368 g imidazole-2-carboxaldehyde to 400 cc DMF and heating the solution to 393 K in an oil bath for 20 minutes. The solution was then cooled to ambient temperature and left to crystallize for 4 days. The collected crystals were washed three times with methanol and air dried at 333 K. The synthesis protocol reported by Zhang et al.

27

was modified in order to synthesize ZIF-71 in this work. Two 60 cc methanol solutions were prepared containing 297 mg zinc (II) acetate and 876 mg 4,5-dichloroimidazole. The two solutions were then mixed and kept without stirring for 24 hours under ambient conditions. The collected crystals were washed three times with methanol and air dried at 333 K. Activation of crystals post air-drying was carried out by degassing in vacuum at 453 K for 24 hours.

5 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 36

Characterization Activated ZIF samples were characterized using powder X-ray diffraction (PXRD), nitrogen physisorption (NP) at 77 K, in situ and ex situ Fourier transform infrared spectroscopy (FTIR) and scanning electron microscopy (SEM). An X’Pert Pro PANalytical X-ray Diffractometer operating with a Cu anode at 45 kV and 40 mA was used for collecting PXRD patterns with a scan time of 10 s/step and a step size of 0.02 degrees 2θ over the range of 2.5-50 degrees 2θ. A BET surface area analyzer (Tristar, Micromeritics) was used to measure surface area and pore volume using NP at 77 K using individually determined pressure ranges.28 SEM measurements were carried out with a LEO 1530 scanning electron microscope (Zeiss Electron Microscopy). Samples were coated with gold by sputtering for 60 seconds under vacuum and a 15 kV accelerating voltage was used for imaging. A Thermo Scientific Nicolet iS50 FT-IR equipped with an iS50 ATR module was used for ex situ FTIR spectra collection. Powder ZIF samples were directly analyzed from 550-4000 cm-1 with 32 scans at a 2 cm-1 resolution.

In situ Diffuse Reflectance Infrared Fourier Transform Spectroscopy The in situ Diffuse Reflectance Infrared Fourier Transform Spectroscopy (DRIFTS) experiments were performed on a FTIR spectrometer (Thermo, Nicolet iS50) equipped with a liquid nitrogen cooled MCT/A detector, a diffuse reflectance accessory (Praying Mantis, Harrick), and a high temperature reaction chamber (HVC, Harrick). The chamber used in NO2 exposure experiments was coated with SilcoNert. KBr was loaded into the chamber before each sample and measured as IR background. Pre-activated ZIF samples were loaded into the sample chamber and reactivated in situ at 383 K under 20 cc min-1 He flow for 3 hours. After cooling to 298 K, the He gas was switched to 20 cc min-1 1000 ppm NO2 with balancing N2. IR spectra was

6 ACS Paragon Plus Environment

Page 7 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

recorded at pre-programmed intervals with 32 scans and 4 cm-1 resolution for the duration of exposure. Dry NO2 Exposure 100 mg powder samples were placed in a fritted 6 mm × 4 mm × 4.5 in. (O.D. x I.D. x L) quartz glass thermal desorption tube (Supelco). The packed bed was placed in a custom-made fixed-bed gas exposure setup. To ensure safe hazardous gas testing, the entire fixed-bed system was housed in a well-ventilated chemical hood, with real-time gas sensors for safety. An upstream pressure gauge was utilized to confirm the absence of detectable pressure drop during gas exposure. Additionally, outlet lines were fed to a 1N NaOH solution to scrub eluted acid gas streams. Packed samples were flushed with ultra-high purity nitrogen (Airgas) at 75 cc min-1 and activated in situ for 2 hours at 453 K. After cooling to ambient conditions, 1000 ppm NO2 gas in balance N2, (Airgas), was passed through the fixed-bed at a flow rate of 75c min-1 for about 2 hours and 25 min (100 ppm-days of NO2 gas exposure). Upon completion of the exposure, the bed was flushed with N2 for 30min and the sample removed for further characterization.

Humid NO2 Exposure Activated samples were exposed to ~20 (ppm) of NO2 in air at 75% relative humidity (R.H.) for 5 days (~100 ppm-days) at 298 K. The NO2 gas was generated from a 400 cc aqueous solution of 0.5 mg/mL NaNO2 at a pH of 4.0 at 318 K in accordance with reported literature.29 Air at 40 cc min-1 was bubbled through the solution and carried humid NO2 gas stream into the exposure unit (Secador mini-desiccator), where the portable PAC 7000 NO2 detector (Dräger) measured the NO2 concentrations at regular intervals. The R.H. was monitored inside the transparent exposure unit using a humidity sensor (Ambient Weather). The NaNO2 solution was

7 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 36

refilled every 24 hours and the custom-made unit was placed inside a fume hood for safe operation. Samples following exposure were re-activated at 453 K for 24 hours in vacuum.

Computational Methods Plane-wave density functional theory (DFT) was used to optimize the experimentally reported structure of ZIF-8.30-33 Calculations were performed in the Vienna Ab-initio Simulation Package (VASP) with projector-augmented wave (PAW) method pseudopotentials34and PerdewBurke-Ernzerhof (PBE) generalized-gradient approximation (GGA) functional35 A conjugate gradient algorithm with a cutoff energy of 480 eV was used to relax atomic positions until all forces were less than 0.05 eV/Å. Same cutoffs and tolerances were used to optimize atomic positions and lattice constants and reciprocal space was sampled only at the Γ-point in all calculations.

Results and Discussion The crystallinity of the three ZIFs before and after exposure to different conditions were characterized by PXRD (Fig. 1). For completeness, the PXRD patterns after exposure to humid air, dry and humid SO2 are reproduced from our previous work.10, 36 Exposure to acid gases is denoted in ppm-days. With 100 ppm-days of exposure to dry NO2, the ZIF-8 and ZIF-90 PXRD patterns show a progressive increase in the background signal, indicating structural degradation. The changes in ZIF-71 are slower, but peak intensities decrease considerably after 1000 ppmdays exposure (Fig. 1 and Fig. S1). This behavior is quite different from exposure to dry SO2, in which all three ZIFs are very stable.10 Under humid NO2 exposures at 75% RH for 100 ppmdays, all three ZIFs exhibit changes in PXRD patterns. In ZIF-8, the changes are less discernable

8 ACS Paragon Plus Environment

Page 9 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

under humid NO2 than similar dry NO2 exposures, indicating slower degradation in humid NO2. ZIF-90, exhibits larger changes under humid NO2 exposure. ZIF-71 transforms to a different crystalline phase (non-porous ZIF-7237), in contrast to its high stability under humid SO2.10, 36 These findings indicate that the interactions of NO2 with ZIFs are very different than those of SO2 or CO2.

Figure 1. PXRD patterns of A) ZIF-8, B) ZIF-90 and C) ZIF-71 after exposure to dry and humid NO2 compared to pre-exposed samples. Figure legends in A and B are the same. *PXRD patterns of ZIFs on exposure to humid air, dry and humid SO2 are reproduced for comparison from previous work.10 The patterns are normalized with respect to the most intense Bragg peak for each ZIF. 9 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 36

Table 1. Textural characteristics of pre-exposed ZIFs and ZIFs after different exposure experiments. The BET surface area is reported as a percentage relative to pre-exposed ZIF-8. * Previously reported data10 # Material undergoes crystal phase change a Average SO2 concentration during dry SO2 isotherm measurements

Exposure Condition

Exposed Time (days)

Conc.

Overall Exposure (ppmdays)

Relative BET SA (%)

Pore Volume (cc/g)

ZIF-8

ZIF-90

ZIF-71

ZIF-8

ZIF-90

ZIF-71

N/A

100

100

100

0.62

0.45

0.37

1.8 ×105

98

99

98

0.6

0.44

0.34

Pre

N/A

N/A

Dry SO2*

0.15

99.8%

Dry NO2

0.1

1000 ppm

100

65

53

99

0.41

0.25

0.36

Dry NO2

1

1000 ppm

1/000

N/A

N/A

45

N/A

N/A

0.17

Humid Air* Humid SO2*

5

0

0

99

98

99

0.61

0.44

0.33

5

20 ppm

100 61

50

97

0.38

0.22

0.34

Humid NO2

5

20 ppm

100

81

3

0#

0.49

0.04

0#

a

Fig. S2 and Table 1 show data from nitrogen physisorption (NP) isotherms at 77 K of the exposed materials. Decreasing NP profiles are observed for ZIF-8 and ZIF-90, with a greater decrease in ZIF-8 upon dry (rather than humid) NO2 exposure. In contrast, ZIF-90 exhibits nearcomplete porosity loss upon humid NO2 exposure. The ZIF-71 profiles exhibit no significant change after 100 ppm-day dry NO2 exposure but decrease considerably after 1000 ppm-days dry exposure. After humid NO2 exposure of ZIF-71, no porosity in NP is observed due to the phase 10 ACS Paragon Plus Environment

Page 11 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

change to nonporous ZIF-72. These PXRD and NP data fully corroborate each other, indicate remarkably different interactions of NO2 with ZIFs than SO2 and CO2, and show that none of the three ZIFs are stable to NO2 exposure under either dry or humid (75% RH) conditions.

In situ FTIR spectroscopy was used to probe chemical bonding changes in each ZIF over a 5-hr exposure to 1000 ppm dry NO2, and the temporal evolution of the several new peaks are shown in the difference spectra (Figs. 2A-2F). Absorbance peaks (marked in black) signify formation of new chemical species, while negative peaks (marked in red) indicate reduced concentration of pre-existing moieties in the ZIF after gas exposure. The raw FTIR spectra are available in Fig. S3. In ZIF-8, reduced intensities observed at 760 and 1146 cm-1 correspond to aromatic C-H bending vibrations while those at 1310, 1430 and 1459 cm-1 correspond to the C-H bending of the methyl group.38 The intensity reduction observed at 990 cm-1 can be attributed to imidazole ring twisting while those at 2930 and 3135 cm-1 are due to the aliphatic and aromatic C-H stretch respectively.38-39 ZIF-90 and ZIF-71 also show reduced intensities for aromatic and aliphatic C-H bending and stretching vibrations. In ZIF-90, the reduction at 1700 cm-1 corresponds to the C=O stretch. In ZIF-71, which has no aliphatic C-H groups, only one major peak reduction is observed at 3130 cm-1 for the aromatic C-H stretch. These findings point to strong interactions/reactions of NO2 with the aromatic and aliphatic C-H groups in the imidazole rings of all three ZIFs. Due to the large number of peaks, more definite attributions to chemical species have been carried out by peak-fitting the FTIR spectra and literature references. Table S2 shows the assignments discussed below, with unambiguously attributed peaks highlighted in green.

11 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 36

The asymmetric NO2 stretch of nitro- groups and nitramines lead to the appearance of peaks between 1500-1600 cm-1, with the nitro- stretch in the lower wavenumber range (15001540 cm-1) and the nitramine stretch at higher wavenumbers.40-42 New FTIR peaks at 1530, 1525 and 1540 cm-1 in ZIF-8, ZIF-90 and ZIF-71 respectively were attributed to nitro- groups, with those at 1560-1570 cm-1 to nitramines. Peaks corresponding to the symmetric stretch of NO2 in the nitro- group were present in the range 1330-1370 cm-1 in ZIF-8 and ZIF-71, but are masked in ZIF-90 by the strong reduction in intensity at 1315 cm-1. We assigned the new peaks at 1031, 1010 and 1020 cm-1 in the three ZIFs to the C-N stretch of the nitro- group, based on the C-N stretch assignment for aromatic nitro- group containing imidazoles such as tinidazole and metronidazole.43-44

Peaks around 930 cm-1 observed in ZIF-8 and ZIF-90 can arise from

additional C-N stretches due to aliphatic nitro- group formation.40 The absence of peaks in this region in ZIF-71, which lacks aliphatic C-H groups, supports this assignment. The N-N stretch of nitramines were assigned to the peak around 960-980 cm-1 in all the three ZIFs, in accordance with literature reports.40, 45 FTIR peaks in the range of 1360-1400 cm-1 can be unambiguously assigned to inorganic nitrate groups bonded to the metal center of the ZIFs.40, 46-47 The peak at ~1380 cm-1 present in all 3 ZIFs has therefore been assigned to the inorganic nitrate group, in accordance with the FTIR spectra of zinc nitrate.48-50 The inorganic nitrite peaks in the 12301280 cm-1 range are difficult to uniquely identify because of overlaps with other functional groups. Various peaks appear in the 1230-1300 cm-1 region in all three ZIFs upon dry NO2 exposure, pointing to formation of multiple species including nitramines, organic nitrates and inorganic nitrites. 40-42, 45, 51-52 Organic nitrates and nitrites have strong stretches in the 1600-1680 cm-1 region and the peaks at ~1620 and ~1670 cm-1 have been assigned to nitrates and nitrites respectively in all three ZIFs.40 The C-O stretches for organic nitrates or nitrites could be present

12 ACS Paragon Plus Environment

Page 13 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

within a wide range and are assigned peaks around ~1100 cm-1, based upon the C-O stretch of ethers and esters.40 Different NO2 deformation vibrations and N-O stretches of multiple species lie in the 650-850 cm-1 range and have been assigned in Table S2. Peaks at 1450-1500 cm-1 result from the formation of nitrosamines (N=O stretch), with the aromatic C-N stretch of nitramines/nitrosamines around 1160 cm-1.40,

53

Broad O-H stretches (hydrogen bonding) are

observed in all three ZIFs around 3300 cm-1. The peak at 1760 cm-1 is likely due to adsorbed N2O454 while the peak around 1700 cm-1 in ZIF-8 and ZIF-71 is likely due to adsorbed nitrous acid.42, 52, 55 Dry NO2 exposure thus leads to a diverse set of new organic and inorganic species, strongly suggesting attack of NO2 on the imidazole linkers as well as the Zn-N coordination bonds. Our analysis above reveals formation of nitro, organic nitrite, nitrosamine, nitramine, and inorganic nitrate groups. The data also suggests the possible presence of organic nitrates and inorganic nitrites, although no peak can be unambiguously associated to these groups.

13 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 36

Figure 2. In situ FTIR difference spectra of ZIFs exposed to 1000 ppm dry NO2 over 5 hours. Decreasing peaks are marked in red. Time intervals of the FTIR spectra are identical in A and B (ZIF-8), C and D (ZIF-90) and E and F (ZIF-71).

14 ACS Paragon Plus Environment

Page 15 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Since the in situ IR system is not compatible with corrosive gases, we next used ex situ ATR FTIR spectroscopy to investigate changes in ZIFs under humid NO2 (Fig. 3 and S5). To confirm qualitative agreement between the two methods, we also measured the ex situ ATR FTIR spectra of dry NO2-exposed ZIFs (Fig. S4) for comparison with the in situ spectra of Fig. S3. The pre-exposed ex situ and in situ FTIR spectra are consistent across the three materials. Specifically, the ZIF-8 spectrum in Figure S4 consists of peaks below 800 cm-1 (out-of-plane bending of the ring), 900-1350 cm-1 (in-plane bending of the imidazole ring), and between 1350 and 1500 cm-1 (vibration of the imidazole ring) respectively.

39, 56

Similar assignments for these

regions can be made in ZIF-71 and ZIF-90, where these imidazole ring bending vibrations dominate the fingerprint region of the spectra and are difficult to uniquely identify. In ZIF-8, peaks at 1146 cm–1 and 1581 cm-1 may be assigned to =C-H bending and C=N stretching respectively.38-39, 56 The symmetric and asymmetric methyl group C-H stretch in ZIF-8 can be observed at 2930 cm-1 and 2980 cm-1 respectively while the aromatic ring C-H stretch, observed at 3135 cm-1, is consistent with literature reports.38-39 In ZIF-71, peaks observed at 665 cm-1 can be attributed to C-Cl vibrations,

57-59

while the characteristic C=O stretch of ZIF-90 is observed

at 1680 cm-1 consistent with literature reports.

32, 60-61

The peaks observed in ex situ FTIR

immediately after dry NO2 exposure (Fig. S4) also match those observed in situ (Fig. S3). Comparison of the ex situ spectra before and after (Fig. S4) reactivation at 180°C under vacuum for 24 hours, shows that the new chemical species formed during the exposure are not completely removed. Hence, a permanent alteration to all the three ZIF structures results even upon dry NO2 exposure.

15 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 36

Figure 3. Ex situ FTIR spectra of ZIFs with identical NO2 dosage under dry and humid (75% RH) conditions. Important peaks discussed in the manuscript are marked in the figure. The FTIR spectra of ZIF-8 and ZIF-90 upon dry and humid NO2 exposure are in general agreement (Fig. 3 and Fig. S5), although decreases in absorbance upon humid NO2 exposures are less intense than the equivalent dry exposures (e.g., the

decrease in imidazole ring

absorbances such as 1590 cm-1 which likely corresponds to the C=N imidazole stretch).39 However, the FTIR spectra of ZIF-71 on humid and dry NO2 exposures are significantly different. In contrast with ZIF-8 and ZIF-90, the imidazole ring vibration absorbances increase 16 ACS Paragon Plus Environment

Page 17 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

upon humid NO2 exposure. This is due to phase transformation to the dense, non-porous ZIF-72 which thus has a higher concentration of chemical bonds. A peak at 1385 cm-1 corresponding to inorganic nitrate is also observed on 100 ppm-day humid NO2 exposure in ZIF-71. Additionally, the morphological changes in the ZIFs upon NO2 exposure were examined by SEM (Fig. S6). ZIF-8 shows significant changes upon dry NO2 exposure with many irregular fragments and a general absence of the crystal facets that were present before exposure. ZIF-71 shows no significant morphological changes upon 1000 ppm-days dry NO2 exposure even though it lost ~50% BET surface area. In contrast, after a 100 ppm-days humid NO2 exposure, the ZIF-71 crystals are etched. The color of each exposed ZIF powder sample also changes (Fig. S7).

Humid NO2 exposure has very different effects on the three ZIFs. In the hydrophobic ZIF-8, the presence of humidity slows the degradation process in relation to dry NO 2, a surprising behavior also reported for the hydrophobic MOF UiO-66.18 In the hydrophilic ZIF-90, humidity accelerates framework degradation. In the hydrophobic ZIF-71, the synergy of humidity and NO2 drives a phase change to nonporous ZIF-72, an entirely different outcome from the slow degradation route observed under prolonged dry NO2. Based upon consideration of all the foregoing results as well as literature information on NO2 reactivity, we propose the mechanisms depicted in Fig. 4 for dry NO2 degradation of ZIFs. NO2 addition to unsaturated bonds, hydrogen abstraction, and radical dimerization, in addition to its strong oxidizing action, are well known.62-64

NO2 reacts with unsaturated organics in the gas phase at ambient

temperature forming nitrates or nitrites, with the hydrogen abstraction route preferred at lower concentrations (~1000 ppm).65-66 Nitrous acid (HONO), formed from hydrogen abstraction, can be produced with ~100% yield in reactions of NO2 with dry soot at ambient conditions.67

17 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 36

Figure 4. Reactants and product species generated during degradation of ZIFs under dry NO2 exposure: (A) Stoichiometrically balanced reactions of ZIF-8 (i), ZIF-90 (ii) and ZIF-71 (iii) are individually shown, while (B) and (C) are valid for any of the ZIF linkers (having general functional groups R1, R2 and R3 at the 2, 4, and 5-positions).

18 ACS Paragon Plus Environment

Page 19 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Formation of nitrites, nitrates, nitro-, and nitramine groups have been reported during gas-phase reactive adsorption of NO2 on a variety of carbonaceous materials at ambient temperature.41, 51, 62, 68-71

The presence of imidazoles in ZIFs creates the strong possibility of similar mechanisms. The

strong decrease in aromatic and aliphatic C-H stretching and bending vibrations in all three ZIFs upon dry NO2 exposure, and the presence of adsorbed HONO (1700 cm-1 stretch), are consistent with the free radical H-abstraction mechanism. While the direct addition of NO2 to the aromatic ring is possible, the above evidence as well as the low NO2 concentration point towards a dominant H-abstraction mechanism (Fig. 4a). It proceeds with NO2 abstracting H from the linker to form HONO, allowing additional NO2 free radicals to react with the imidazole radical that was created, and thus forming nitro, organic nitrite or nitrate species.62, 65 HONO can dissociate to create NO and a reactive hydroxyl radical67, 72, which can initiate a chain reaction with other free radicals such as NO2 or carry out H-abstraction reactions to form water. The decreased crystallinity of the ZIFs after dry NO2 exposure, and the presence of inorganic nitrites and nitrates, are consistent with Zn-N bond cleavage. HONO produced by H-abstraction attacks the Zn-N bonds (Fig. 4b) protonating the imidazole N and forming an inorganic nitrite, which is then oxidized73 by NO2 to form an inorganic nitrate. NO2 or HONO react with the protonated imidazole74-77 to form nitramines and nitrosamines. In summary, the free radical H-abstraction mechanism explains the formation of organic N-containing groups (observed in the FTIR spectra) leading to functionalized imidazole linkers (rather than degradation of the ZIF per se); whereas the observed formation of nitrosamines, nitramines and inorganic nitrites/nitrates results from secondary reactions with HONO produced by H-abstraction and is responsible for the Zn-N bond cleavages and degradation of the ZIF structure. In addition, NO2 is a strong oxidizing agent that has been observed to oxidize ionic salts in the dark at ppm concentrations to form inorganic

19 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 36

nitrites, which can be further oxidized by NO2 to nitrates with NO evolution.73,

78-79

Reactive

adsorption of NO2 on metals (including Zn) and metal oxides is known to proceed at ambient temperature via disproportionation of two NO2 molecules or the N2O4 dimer to form a surface nitrate and evolve NO.80-83 Thus, the metal centers in ZIFs are a third route for NO2 interactions, also leading to Zn-N bond cleavage (Fig. 4c).

The finding that ZIF-71 is the most stable of the three ZIFs under dry NO2 exposure provides further evidence regarding the dominant mechanism among those in Fig. 4. The ZIF-71 linker has only a single (aromatic) C-H bond, unlike ZIF-8 or ZIF-90 linkers which have multiple aromatic (two each) and aliphatic (three and one, respectively) C-H bonds. This strongly suggests that the dominant ZIF degradation mechanism is H-abstraction from the linkers, leading to secondary reactions of the resulting acidic species that are responsible for Zn-N bond cleavage. Direct attack on Zn sites by reactive dry NO2 adsorption cannot be ruled out, but the observed speciation patterns and the significant difference in reactivity of the three ZIFs suggest that it plays a minor role at best. We obtained additional insight through periodic DFT calculations to compute defect formation energies84-85 for specific reactions of ZIF-8 (SOD topology) in NO2 environments, by subtracting the total energy of the products from the reactants

(Table 2, Reactions

1-6).

Negative defect

formation

energies imply a

thermodynamically favorable degradation reaction. These results agree with our proposed mechanism. The H-abstraction by dry NO2 in ZIF-8 forming nitro or nitrite groups along with HONO are all calculated to be very thermodynamically favorable (Reactions 1-3). Attack on the ZIF-8 structure

by

HONO or HNO3 forming inorganic nitrites/nitrates is also favorable

(Reactions 4-5) albeit with considerably smaller defect formation energies.

20 ACS Paragon Plus Environment

Page 21 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Table 2. Defect formation energies (eV) of reactions involving ZIF-8 computed using DFT. #

Reaction Scheme

Defect Energy (eV)

1

-1.95

2

-1.73

3

-1.69

4

-0.10

5

-0.29

6

1.07

21 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 36

Interestingly, direct oxidation of Zn by reactive adsorption of NO2 (Reaction 6) is very unfavorable. Unfortunately, the high computational costs associated with the much larger unit cell of the RHO topology of ZIF-71 rendered calculations for this material infeasible. All the discussion above focused on dry NO2 exposure. Under humid NO2, the degradation mechanism has an additional contribution. The reaction of NO2 and water vapor (in the dark and irrespective of O2 availability) occurs on any reactor surface walls (pyrex glass, quartz, teflon) as well as other materials (silica, activated carbon, alumina, glass), leading to formation of gaseous nitrous acid and adsorbed nitric acid.86-90 According to the proposed mechanism90-91, NO2 (or N2O4) is absorbed into an aqueous surface film, the HONO product either desorbs into the gas phase or reacts further to form NO, and the HNO3 product remains adsorbed on the surface. Our investigation indicates that humidity impedes the reaction of NO2 with hydrophobic ZIFs such as ZIF-8, most likely via competitive adsorption of NO2 and water. In hydrophilic ZIF-90, dissolution of NO2 in the water-filled pores leads to HNO3 formation and aids in faster degradation of the ZIF-90 structure, as evinced by an enhanced reduction of imidazole ring stretches (Fig. 3). Reductions in intensity (relative to dry exposures) of the peaks at ~1290 cm-1 upon humid NO2 exposure of ZIF-90 suggest that the sites of interaction of HNO3 with the ZIF90 structure differ from those of dry NO2 whose action is impeded by water. The surprising phase transformation of ZIF-71 to ZIF-72 under humid NO2 points to bond cleavages induced by small amounts of surface-adsorbed HNO3 formed under humid NO2 exposure, a hypothesis that is supported by the presence of inorganic nitrates in its FTIR spectra and the absence of nitro-, nitramine, organic nitrites and nitrate groups (Fig. 3c). Similar actions of adsorbed HNO3 are also expected in hydrophobic ZIF-8, but unlike ZIF-71 it does not undergo a phase transformation and instead degrades slowly.

22 ACS Paragon Plus Environment

Page 23 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Conclusions We have conducted the first systematic investigation of the effects of dry and humid NO2 gas on three representative ZIFs, chosen carefully based upon our previous work. All the ZIFs investigated are shown to be unstable under prolonged dry or humid NO2 exposures. The hydrophobic ZIF-8 degrades faster under dry NO2 exposure than the hydrophilic ZIF-90, but in contrast degrades more slowly than ZIF-90 under humid NO2 exposure. Hydrophobic ZIF-71 shows the slowest degradation under dry NO2 exposure but undergoes a structural transformation to the nonporous but chemically identical ZIF-72 on prolonged humid NO2 exposure. These findings are in strong contrast with the interactions of dry or humid SO2 with ZIFs.10,

36

The

presence of various nitro-, organic and inorganic nitrites and nitrates, nitrosamines and nitramines were supported by FTIR measurements, which revealed the mechanisms of ZIF degradation by dry and humid NO2. Our investigation strongly suggests the free radical Habstraction by NO2 (both dry and humid) and HNO3 formation (humid) as major interaction routes, with possible minor contributions from NO2 disproportionation on Zn metal centers. DFT simulations of defect energy formation support this mechanistic picture. The much higher reactivity of dry NO2 over dry SO2 can be attributed to its free radical nature. We envision similar mechanisms of NO2 attack on other MOFs, which make NO2 a much more potent hindrance to widespread use of MOFs in acid gas-related applications and would require new approaches to stabilize ZIF/MOF materials.

23 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 36

Associated Content Supporting Information Characteristics of the 3 single-linker ZIFs, PXRD spectra showing evolution of the (110) peak of ZIF-71 on prolonged dry NO2 exposure, textural properties of the ZIFs on NO2 exposure, raw insitu DRIFTS spectra for each ZIF, list of peak assignments from FTIR spectroscopy, ex-situ ATR FTIR spectra of the three ZIFs before and after activation on dry NO2 exposure, ex-situ ATR FTIR spectra of the three reactivated ZIFs under dry and humid NO2 exposure, SEM images showing changes in morphology of the 3 ZIFs and the color changes observed on dry and humid NO2 exposure are included in the supplementary information. This material is available free of charge via the Internet at http://pubs.acs.org.

Author Information Corresponding Authors *[email protected], [email protected]

Author Contributions The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript.

Notes The authors declare no competing financial interest.

24 ACS Paragon Plus Environment

Page 25 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Acknowledgments This work was supported as part of UNCAGE-ME, an Energy Frontier Research Center funded by the U.S. Department of Energy, Office of Science, Basic Energy Sciences under Award # DESC0012577.

References (1)

Levasseur, B.; Petit, C.; Bandosz, T. J. Reactive Adsorption of NO2 on Copper-Based

Metal-Organic Framework and Graphite Oxide/Metal-Organic Framework Composites. ACS Appl. Mater. Interfaces 2010, 2, 3606-3613. (2)

Nriagu, J. O.; Simmons, M. S. Environmental Oxidants; John Wiley & Sons, Inc.: New

York, U.S.A.; 1994. (3)

Rezaei, F.; Rownaghi, A. A.; Monjezi, S.; Lively, R. P.; Jones, C. W. SOx/NOx Removal

from Flue Gas Streams by Solid Adsorbents: A Review of Current Challenges and Future Directions. Energy Fuels 2015, 29, 5467-5486. (4)

U.S.E.P.A. Air Pollution Control Technology Fact Sheet: Selective Catalytic Reduction.

https://www3.epa.gov/ttncatc1/dir1/fscr.pdf, 2018 (accessed December 18, 2018). (5)

Sumida, K.; Rogow, D. L.; Mason, J. A.; McDonald, T. M.; Bloch, E. D.; Herm, Z. R.;

Bae, T.-H.; Long, J. R. Carbon Dioxide Capture in Metal–Organic Frameworks. Chem. Rev. 2012, 112, 724-781. (6)

Banerjee, R.; Furukawa, H.; Britt, D.; Knobler, C.; O'Keeffe, M.; Yaghi, O. M. Control

of Pore Size and Functionality in Isoreticular Zeolitic Imidazolate Frameworks and Their Carbon Dioxide Selective Capture Properties. J.Am.Chem.Soc. 2009, 131, 3875-3877.

25 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(7)

Page 26 of 36

Phan, A.; Doonan, C. J.; Uribe-Romo, F. J.; Knobler, C. B.; O'Keeffe, M.; Yaghi, O. M.

Synthesis, Structure, and Carbon Dioxide Capture Properties of Zeolitic Imidazolate Frameworks. Acc. Chem. Res. 2010, 43, 58-67. (8)

Pimentel, B. R.; Parulkar, A.; Zhou, E.-k.; Brunelli, N. A.; Lively, R. P. Zeolitic

Imidazolate Frameworks: Next-Generation Materials for Energy-Efficient Gas Separations. Chemsuschem 2014, 7, 3202-3240. (9)

Walton, K. S.; Sholl, D. S. Research Challenges in Avoiding “Showstoppers” in

Developing Materials for Large-Scale Energy Applications. Joule 2017, 1, 208-211. (10)

Bhattacharyya, S.; Han, R.; Kim, W.-G.; Chiang, Y.; Jayachandrababu, K. C.;

Hungerford, J. T.; Dutzer, M. R.; Ma, C.; Walton, K. S.; Sholl, D. S., et al. Acid Gas Stability of Zeolitic Imidazolate Frameworks: Generalized Kinetic and Thermodynamic Characteristics. Chem. Mater. 2018, 30, 4089-4101. (11)

Jiao, Y.; Liu, Y.; Zhu, G.; Hungerford, J. T.; Bhattacharyya, S.; Lively, R. P.; Sholl, D.

S.; Walton, K. S. Heat-Treatment of Defective UiO-66 from Modulated Synthesis: Adsorption and Stability Studies. J. Phys. Chem. C 2017, 121, 23471-23479. (12)

Mounfield, W. P.; Han, C.; Pang, S. H.; Tumuluri, U.; Jiao, Y.; Bhattacharyya, S.;

Dutzer, M. R.; Nair, S.; Wu, Z.; Lively, R. P., et al. Synergistic Effects of Water and SO2 on Degradation of MIL-125 in the Presence of Acid Gases. J. Phys. Chem. C 2016, 120, 2723027240. (13)

Bobbitt, N. S.; Mendonca, M. L.; Howarth, A. J.; Islamoglu, T.; Hupp, J. T.; Farha, O.

K.; Snurr, R. Q. Metal-Organic Frameworks for the Removal of Toxic Industrial Chemicals and Chemical Warfare Agents. Chem. Soc. Rev. 2017, 46, 3357-3385.

26 ACS Paragon Plus Environment

Page 27 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(14)

DeCoste, J. B.; Peterson, G. W. Metal–Organic Frameworks for Air Purification of Toxic

Chemicals. Chem. Rev. 2014, 114, 5695-5727. (15)

Petit, C.; Levasseur, B.; Mendoza, B.; Bandosz, T. J. Reactive Adsorption of Acidic

Gases on MOF/Graphite Oxide Composites. Microporous Mesoporous Mater. 2012, 154, 107112. (16)

DeCoste, J. B.; Demasky, T. J.; Katz, M. J.; Farha, O. K.; Hupp, J. T. A UiO-66

Analogue with Uncoordinated Carboxylic Acids for the Broad-Spectrum Removal of Toxic Chemicals. New J. Chem. 2015, 39, 2396-2399. (17)

Ebrahim, A. M.; Bandosz, T. J. Effect of Amine Modification on the Properties of

Zirconium–Carboxylic Acid Based Materials and Their Applications as NO2 Adsorbents at Ambient Conditions. Microporous Mesoporous Mater. 2014, 188, 149-162. (18)

Ebrahim, A. M.; Levasseur, B.; Bandosz, T. J. Interactions of NO2 with Zr-Based MOF:

Effects of the Size of Organic Linkers on NO2 Adsorption at Ambient Conditions. Langmuir 2013, 29, 168-174. (19)

Peterson, G. W.; Mahle, J. J.; DeCoste, J. B.; Gordon, W. O.; Rossin, J. A. Extraordinary

NO2 Removal by the Metal–Organic Framework UiO-66-NH2. Angew. Chem., Int. Ed. 2016, 55, 6235-6238. (20)

Han, S.; Huang, Y.; Watanabe, T.; Dai, Y.; Walton, K. S.; Nair, S.; Sholl, D. S.;

Meredith, J. C. High-Throughput Screening of Metal-Organic Frameworks for CO2 Separation. ACS Comb. Sci. 2012, 14, 263-267. (21)

Burtch, N. C.; Jasuja, H.; Walton, K. S. Water Stability and Adsorption in Metal-Organic

Frameworks. Chem. Rev. 2014, 114, 10575-10612.

27 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(22)

Page 28 of 36

Yang, L.-M.; Vajeeston, P.; Ravindran, P.; Fjellvåg, H.; Tilset, M. Revisiting Isoreticular

Mofs of Alkaline Earth Metals: A Comprehensive Study on Phase Stability, Electronic Structure, Chemical Bonding, and Optical Properties of a–IRMOF-1 (a = Be, Mg, Ca, Sr, Ba). Phys. Chem. Chem. Phys. 2011, 13, 10191-10203. (23)

Yang, L.-M.; Dornfeld, M.; Hui, P.-M.; Frauenheim, T.; Ganz, E. Ten New Predicted

Covalent Organic Frameworks with Strong Optical Response in the Visible and near Infrared. J.Chem.Phys. 2015, 142, 244706. (24)

Chung, Y. G.; Camp, J.; Haranczyk, M.; Sikora, B. J.; Bury, W.; Krungleviciute, V.;

Yildirim, T.; Farha, O. K.; Sholl, D. S.; Snurr, R. Q. Computation-Ready, Experimental MetalOrganic Frameworks: A Tool to Enable High-Throughput Screening of Nanoporous Crystals. Chem. Mater. 2014, 26, 6185-6192. (25)

Nazarian, D.; Camp, J. S.; Sholl, D. S. A Comprehensive Set of High-Quality Point

Charges for Simulations of Metal-Organic Frameworks. Chem. Mater. 2016, 28, 785-793. (26)

Gee, J. A.; Chung, J.; Nair, S.; Sholl, D. S. Adsorption and Diffusion of Small Alcohols

in Zeolitic Imidazolate Frameworks ZIF-8 and ZIF-90. J. Phys. Chem. C 2013, 117, 3169-3176. (27)

Zhang, K.; Lively, R. P.; Dose, M. E.; Brown, A. J.; Zhang, C.; Chung, J.; Nair, S.;

Koros, W. J.; Chance, R. R. Alcohol and Water Adsorption in Zeolitic Imidazolate Frameworks. Chem. Commun. 2013, 49, 3245-3247. (28)

Walton, K. S.; Snurr, R. Q. Applicability of the Bet Method for Determining Surface

Areas of Microporous Metal-Organic Frameworks. J.Am.Chem.Soc. 2007, 129, 8552-8556. (29)

Hashimoto, Y.; Tanaka, S. A New Method of Generation of Gases at Parts Per Million

Levels for Preparation of Standard Gases. Environ. Sci. Technol. 1980, 14, 413-416.

28 ACS Paragon Plus Environment

Page 29 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(30)

Park, K. S.; Ni, Z.; Cote, A. P.; Choi, J. Y.; Huang, R. D.; Uribe-Romo, F. J.; Chae, H.

K.; O'Keeffe, M.; Yaghi, O. M. Exceptional Chemical and Thermal Stability of Zeolitic Imidazolate Frameworks. Proc. Natl. Acad. Sci. U. S. A. 2006, 103, 10186-10191. (31)

Banerjee, R.; Phan, A.; Wang, B.; Knobler, C.; Furukawa, H.; O'Keeffe, M.; Yaghi, O.

M. High-Throughput Synthesis of Zeolitic Imidazolate Frameworks and Application to CO2 Capture. Science 2008, 319, 939-943. (32)

Morris, W.; Doonan, C. J.; Furukawa, H.; Banerjee, R.; Yaghi, O. M. Crystals as

Molecules: Postsynthesis Covalent Functionalization of Zeolitic Imidazolate Frameworks. J.Am.Chem.Soc. 2008, 130, 12626-12627. (33)

First, E. L.; Floudas, C. A. Mofomics: Computational Pore Characterization of Metal–

Organic Frameworks. Microporous Mesoporous Mater. 2013, 165, 32-39. (34)

Blöchl, P. E. Projector Augmented-Wave Method. Phys. Rev. B 1994, 50, 17953-17979.

(35)

Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made

Simple. Phys. Rev. Lett. 1996, 77, 3865-3868. (36)

Bhattacharyya, S.; Pang, S. H.; Dutzer, M. R.; Lively, R. P.; Walton, K. S.; Sholl, D. S.;

Nair, S. Interactions of SO2-Containing Acid Gases with ZIF-8: Structural Changes and Mechanistic Investigations. J. Phys. Chem. C 2016, 120, 27221-27229. (37)

Schweinefuss, M. E.; Springer, S.; Baburin, I. A.; Hikov, T.; Huber, K.; Leoni, S.;

Wiebcke, M. Zeolitic Imidazolate Framework-71 Nanocrystals and a Novel SOD-Type Polymorph: Solution Mediated Phase Transformations, Phase Selection Via Coordination Modulation and a Density Functional Theory Derived Energy Landscape. Dalton Trans. 2014, 43, 3528-3536.

29 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(38)

Page 30 of 36

Cheng, P.; Hu, Y. H. H2O-Functionalized Zeolitic Zn(2-Methylimidazole)2 Framework

(ZIF-8) for H2 Storage. J. Phys. Chem. C 2014, 118, 21866-21872. (39)

Hu, Y.; Kazemian, H.; Rohani, S.; Huang, Y.; Song, Y. In Situ High Pressure Study of

ZIF-8 by FTIR Spectroscopy. Chem. Commun. 2011, 47, 12694-12696. (40)

Socrates, G. Infrared and Raman Characteristic Group Frequencies; John Wiley & Sons,

Ltd.: New York, U.S.A.; 2001. (41)

Akhter, M. S.; Chughtai, A. R.; Smith, D. M. Reaction of Hexane Soot with Nitrogen

Dioxide/Nitrogen Oxide (N2O4). J. Phys. Chem. 1984, 88, 5334-5342. (42)

Al-Abadleh, H. A.; Grassian, V. H. Heterogeneous Reaction of NO2 on Hexane Soot: A

Knudsen Cell and FT-IR Study. J. Phys. Chem. A 2000, 104, 11926-11933. (43)

Azambre, B.; Collura, S.; Trichard, J. M.; Weber, J. V. Nature and Thermal Stability of

Adsorbed Intermediates Formed During the Reaction of Diesel Soot with Nitrogen Dioxide. Appl. Surf. Sci. 2006, 253, 2296-2303. (44)

Mahendra, K. T.; Shrikant, P.; Harish, S.; Khemraj, B.; Snehasis, J. Spectroscopic

Characterization of Biofield Treated Metronidazole and Tinidazole. Med. Chem. 2015, 5, 340344. (45)

Piskorz, M.; Urbanski, T. Ultraviolet and Infrared Spectra of Some Nitramines Bull.

Acad. Pol. Sci., Ser. Sci. Chim. 1963, 11, 615-624. (46)

W.L.Ross. The Infrared Absorption Spectra of Nitrogen Dioxide. Masters Dissertation,

The University of British Columbia, Canada, 1950. (47)

Brooker, M. H. Raman and I.R. Spectra of Zinc, Cadmium and Calcium Nitrate: A Study

of the Low Temperature Phase Transitions in Calcium Nitrate. Spectroc. Acta Pt. A-Molec. Spectr. 1976, 32, 369-377.

30 ACS Paragon Plus Environment

Page 31 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(48)

Li, P.; Xu, Z. P.; Hampton, M. A.; Vu, D. T.; Huang, L.; Rudolph, V.; Nguyen, A. V.

Control Preparation of Zinc Hydroxide Nitrate Nanocrystals and Examination of the Chemical and Structural Stability. J. Phys. Chem. C 2012, 116, 10325-10332. (49)

Wahab, R.; Ansari, S. G.; Seo, H.-K.; Kim, Y. S.; Suh, E.-K.; Shin, H.-S. Low

Temperature Synthesis and Characterization of Rosette-Like Nanostructures of ZnO Using Solution Process. Solid State Sci. 2009, 11, 439-443. (50)

Wahab, R.; Kim, Y.-S.; Shin, H.-S. Synthesis, Characterization and Effect of pH

Variation on Zinc Oxide Nanostructures. Mater. Trans. 2009, 50, 2092-2097. (51)

Kirchner, U.; Scheer, V.; Vogt, R. FTIR Spectroscopic Investigation of the Mechanism

and Kinetics of the Heterogeneous Reactions of NO2 and HNO3 with Soot. J. Phys. Chem. A 2000, 104, 8908-8915. (52)

Zawadzki, J.; Wisniewski, M.; Skowronska, K. Heterogeneous Reactions of NO2 and

NO-O2 on the Surface of Carbons. Carbon 2003, 41, 235-246. (53)

Piskorz, M.; Urbanski, T. Ultraviolet and Infrared Spectra of Some Nitrosamines Bull.

Acad. Pol. Sci., Ser. Sci. Chim. 1963, 11, 607-613. (54)

Bibart, C. H.; Ewing, G. E. Vibrational-Spectrum, Torsional Potential, and Binding of

Gaseous N2O4. J.Chem.Phys. 1974, 61, 1284-1292. (55)

Kagann, R. H.; Maki, A. G. Infrared Absorption Intensities of Nitrous Acid (HONO)

Fundamental Bands. J. Quant. Spectrosc. Radiat. Transfer 1983, 30, 37-44. (56)

Barbosa, P.; Rosero-Navarro, N. C.; Shi, F.-N.; Figueiredo, F. M. L. Protonic

Conductivity of Nanocrystalline Zeolitic Imidazolate Framework 8. Electrochim. Acta 2015, 153, 19-27.

31 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(57)

Page 32 of 36

Liu, S.; Liu, G.; Zhao, X.; Jin, W. Hydrophobic-ZIF-71 Filled PEBA Mixed Matrix

Membranes for Recovery of Biobutanol Via Pervaporation. J. Membr. Sci. 2013, 446, 181-188. (58)

Springer, S.; Satalov, A.; Lippke, J.; Wiebcke, M. Nanocrystals and Nanomaterials of

Isoreticular Zeolitic Imidazolate Frameworks. Microporous Mesoporous Mater. 2015, 216, 161170. (59)

Japip, S.; Wang, H.; Xiao, Y.; Shung Chung, T. Highly Permeable Zeolitic Imidazolate

Framework (ZIF)-71 Nano-Particles Enhanced Polyimide Membranes for Gas Separation. J. Membr. Sci. 2014, 467, 162-174. (60)

Liu, C.; Liu, Q.; Huang, A. A Superhydrophobic Zeolitic Imidazolate Framework (ZIF-

90) with High Steam Stability for Efficient Recovery of Bioalcohols. Chem. Commun. 2016, 52, 3400-3402. (61)

Yang, T.; Chung, T.-S. Room-Temperature Synthesis of ZIF-90 Nanocrystals and the

Derived Nano-Composite Membranes for Hydrogen Separation. J. Mater. Chem. A 2013, 1, 6081-6090. (62)

Gray, P.; Yoffe, A. D. The Reactivity and Structure of Nitrogen Dioxide. Chem. Rev.

1955, 55, 1069-1154. (63)

Bernard, F.; Cazaunau, M.; Mu, Y.; Wang, X.; Daële, V.; Chen, J.; Mellouki, A. Reaction

of NO2 with Selected Conjugated Alkenes. J. Phys. Chem. A 2013, 117, 14132-14140. (64)

Kaupp, G.; Schmeyers, J. Gas/Solid Reactions with Nitrogen Dioxide. J. Org. Chem

1995, 60, 5494-5503. (65)

Pryor, W. A.; Lightsey, J. W.; Church, D. F. Reaction of Nitrogen-Dioxide with Alkenes

and Poly-Unsaturated Fatty Acids - Addition and Hydrogen Abstraction Mechanisms. J.Am.Chem.Soc. 1982, 104, 6685-6692.

32 ACS Paragon Plus Environment

Page 33 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(66)

Giamalva, D. H.; Kenion, G. B.; Church, D. F.; Pryor, W. A. Rates and Mechanisms of

Reactions of Nitrogen Dioxide with Alkenes in Solution. J.Am.Chem.Soc. 1987, 109, 7059-7063. (67)

Aubin, D. G.; Abbatt, J. P. D. Interaction of NO2 with Hydrocarbon Soot:  Focus on

HONO Yield, Surface Modification, and Mechanism. J. Phys. Chem. A 2007, 111, 6263-6273. (68)

Davydov, E. Y.; Gaponova, I. S.; Pokholok, T. V.; Vorotnikov, A. P.; Pariiskii, G. B.;

Zaikov, G. E. Features of Mechanism of Free Radical Initiation in Polymers under Exposure to Nitrogen Oxides. Polym. Adv. Technol. 2008, 19, 475-484. (69)

Gao, X.; Liu, S.; Zhang, Y.; Luo, Z.; Ni, M.; Cen, K. Adsorption and Reduction of NO2

over Activated Carbon at Low Temperature. Fuel Process. Technol. 2011, 92, 139-146. (70)

Jeguirim, M.; Tschamber, V.; Brilhac, J. F.; Ehrburger, P. Interaction Mechanism of NO2

with Carbon Black: Effect of Surface Oxygen Complexes. J. Anal. Appl. Pyrolysis 2004, 72, 171-181. (71)

Zhang, W.-J.; Bagreev, A.; Rasouli, F. Reaction of NO2 with Activated Carbon at

Ambient Temperature. Ind. Eng. Chem. Res. 2008, 47, 4358-4362. (72)

Akhter, M. S.; Chughtai, A. R.; Smith, D. M. The Structure of Hexane Soot II: Extraction

Studies. Appl. Spectrosc. 1985, 39, 154-167. (73)

O'Neill, E.; Hinrichs, R. Z. Production of Molecular Iodine from the Heterogeneous

Reaction of Nitrogen Dioxide with Solid Potassium Iodide. J. Geophys. Res.: Atmos. 2011, 116, 1-9. (74)

Challis, B. C.; Kyrtopoulos, S. A. The Chemistry of Nitroso-Compounds. Part 11.

Nitrosation of Amines by the Two-Phase Interaction of Amines in Solution with Gaseous Oxides of Nitrogen. J. Chem. Soc., Perkin Trans. 1 1979, 299-304.

33 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(75)

Page 34 of 36

Turney, T. A.; Wright, G. A. Nitrous Acid and Nitrosation. Chem. Rev. 1959, 59, 497-

513. (76)

Zhao, Y.-L.; Garrison, S. L.; Gonzalez, C.; Thweatt, W. D.; Marquez, M. N-Nitrosation

of Amines by NO2 and NO:  A Theoretical Study. J. Phys. Chem. A 2007, 111, 2200-2205. (77)

Grosjean, D. Atmospheric Chemistry of Toxic Contaminants. 6. Nitrosamines: Dialkyl

Nitrosamines and Nitrosomorpholine. J. Air Waste Manage. Assoc. 1991, 41, 306-311. (78)

Li, H.; Zhu, T.; Ding, J.; Chen, Q.; Xu, B. Heterogeneous Reaction of NO2 on the Surface

of NaCl Particles. Sci. China, Ser. B: Chem. 2006, 49, 371-378. (79)

Vogt, R.; Finlayson-Pitts, B. J. A Diffuse Reflectance Infrared Fourier Transform

Spectroscopic Study of the Surface Reaction of NaCl with Gaseous NO2 and HNO3. J. Phys. Chem. 1994, 98, 3747-3755. (80)

Despres, J.; Koebel, M.; Krocher, O.; Elsener, M.; Wokaun, A. Storage of NO2 on

BaO/TiO2 and the Influence of NO. Appl. Catal. B-Environ. 2003, 43, 389-395. (81)

Grossale, A.; Nova, I.; Tronconi, E. Role of Nitrate Species in the "NO2-SCR"

Mechanism over a Commercial Fe-Zeolite Catalyst for SCR Mobile Applications. Catal. Lett. 2009, 130, 525-531. (82)

Rodriguez, J. A.; Jirsak, T.; Dvorak, J.; Sambasivan, S.; Fischer, D. Reaction of NO2 with

Zn and ZnO:  Photoemission, Xanes, and Density Functional Studies on the Formation of NO3. J. Phys. Chem. B 2000, 104, 319-328. (83)

Tournas, A. D.; Potts, A. W. Surface Reactions of NO2 with Machined Zn, Cd, Cu and

Ag Surfaces Studied by He I/II Photoelectron Spectroscopy. J. Electron Spectrosc. Relat. Phenom. 1993, 66, 167-174.

34 ACS Paragon Plus Environment

Page 35 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(84)

Han, C.; Zhang, C.; Tymińska, N.; Schmidt, J. R.; Sholl, D. S. Insights into the Stability

of Zeolitic Imidazolate Frameworks in Humid Acidic Environments from First-Principles Calculations. J. Phys. Chem. C 2018, 122, 4339-4348. (85)

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. (86)

Febo, A.; Perrino, C. Prediction and Experimental Evidence for High Air Concentration

of Nitrous Acid in Indoor Environments. Atmos. Environ. Part A 1991, 25, 1055-1061. (87)

Kleffmann, J.; Becker, K. H.; Wiesen, P. Heterogeneous NO2 Conversion Processes on

Acid Surfaces: Possible Atmospheric Implications. Atmos. Environ. 1998, 32, 2721-2729. (88)

Sakamaki, F.; Hatakeyama, S.; Akimoto, H. Formation of Nitrous Acid and Nitric Oxide

in the Heterogeneous Dark Reaction of Nitrogen Dioxide and Water Vapor in a Smog Chamber. Int. J. Chem. Kinet. 1983, 15, 1013-1029. (89)

Svensson, R.; Ljungström, E.; Lindqvist, O. Kinetics of the Reaction between Nitrogen

Dioxide and Water Vapour. Atmos. Environ. 1987, 21, 1529-1539. (90)

Finlayson-Pitts, B. J.; Wingen, L. M.; Sumner, A. L.; Syomin, D.; Ramazan, K. A. The

Heterogeneous Hydrolysis of NO2 in Laboratory Systems and in Outdoor and Indoor Atmospheres: An Integrated Mechanism. Phys. Chem. Chem. Phys. 2003, 5, 223-242. (91)

Ramazan, K. A.; Syomin, D.; Finlayson-Pitts, B. J. The Photochemical Production of

HONO During the Heterogeneous Hydrolysis of NO2. Phys. Chem. Chem. Phys. 2004, 6, 38363843.

35 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 36 of 36

Table of Contents Graphic

36 ACS Paragon Plus Environment