Understanding The Fascinating Origins of CO2 ... - ACS Publications

Jul 27, 2016 - Shoushun Chen, Bryan E.G. Lucier, Paul D. Boyle, and Yining Huang*. Department of Chemistry, University of Western Ontario, London, ...
1 downloads 0 Views 3MB Size
Subscriber access provided by Missouri S&T Library

Article

Understanding The Fascinating Origins of CO2 Adsorption and Dynamics in MOFs Shoushun Chen, Bryan E.G. Lucier, Paul D. Boyle, and Yining Huang Chem. Mater., Just Accepted Manuscript • DOI: 10.1021/acs.chemmater.6b02239 • Publication Date (Web): 27 Jul 2016 Downloaded from http://pubs.acs.org on July 28, 2016

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 free 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 accessible to all readers and 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.

Chemistry of Materials 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 58

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

Chemistry of Materials

Understanding The Fascinating Origins of CO2 Adsorption and Dynamics in MOFs

Shoushun Chen, Bryan E.G. Lucier, Paul D. Boyle, Yining Huang*

Department of Chemistry, University of Western Ontario, London, Ontario Canada, N6A 5B7

* Corresponding author, email: [email protected]

1

ACS Paragon Plus Environment

Chemistry of Materials

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 58

Abstract Metal-organic frameworks (MOFs) have shown great promise for the adsorption and separation of gases, including the greenhouse gas CO2. In order to improve performance and realize practical applications for MOFs as CO2 adsorbents, a deeper understanding of the number and type of CO2 adsorption mechanisms must be unlocked, along with fine details of CO2 motion within MOFs. Using

several

complementary

characterization

methods

is

a

promising

protocol

for

comprehensively investigating the various host-guest interactions between MOFs and CO2. In this work, a combination of solid state NMR (SSNMR) and single crystal X-ray diffraction (SCXRD) has been utilized to reveal both the location and dynamics of adsorbed CO2 within the related PbSDB and CdSDB MOFs, as well as to probe the role of metal centers in CO2 adsorption.

13

C

SSNMR experiments targeting CO2 reveal the number of unique adsorption sites and the types of CO2 dynamics present, as well as their associated motional rates and angles.

111

Cd and

207

Pb

SSNMR methods are used to probe the influence of CO2 adsorption on the MOF metal centers, and also to investigate the possibility of any metal-guest interactions. SCXRD experiments yield the exact locations and occupancies of adsorbed CO2 in both MOFs; by pairing this information with SSNMR data, a comprehensive model of CO2 adsorption and dynamics in PbSDB and CdSDB has been established. Both MOFs share a common adsorption site in the V-shaped “π-pocket” formed by the phenyl rings of an individual V-shaped organic linker, while CdSDB also features an additional π-pocket adsorption site arising from the phenyl rings of two linkers joined by Cd. SCXRD and SSNMR data indicate that CO2 adsorbed at the SDB-based π-pocket in both MOFs exhibits a local rotation or “wobbling” at an individual adsorption site, as well as a non-localized jumping or “hopping” between symmetry-equivalent adsorption sites. The combined analysis of SCXRD and SSNMR data has the potential to yield rich information regarding guest dynamics, 2

ACS Paragon Plus Environment

Page 3 of 58

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

Chemistry of Materials

adsorption locations, and host-guest interactions in many MOFs.

Introduction Increasing atmospheric levels of the greenhouse gas CO21-2 and concomitant global warming poses a grave danger to many species of life on Earth.3-7 The formidable challenge of stabilizing and eventually reducing atmospheric CO2 levels must be urgently addressed. Metal-organic frameworks (MOFs) are a promising class of materials for CO2 capture and storage due to their porous nature, extremely large surface areas, affinity for guest molecules, and variety of tunable properties.8-9 Specific combinations of metal ions with organic linkers gives rise to MOFs which possess permanent porosity and can feature one-, two-, or three-dimensional crystalline channels.10-11 The nearly limitless variations of different metals and linkers have led to the preparation and study of many MOFs over the past two decades,10 with potential applications in fields such as gas storage,12-15 gas and liquid separations,16-18 catalysis,19-20 chemical sensors,21-23 and drug delivery.24-26 In particular, there are numerous MOFs that are capable of CO2 adsorption and capture.9, 27-29 At the present time, MOFs cannot selectively adsorb sufficient quantities of CO2 to be employed in automobiles and industry under practical conditions;30 however, many exciting avenues are being explored to address specific obstacles. To improve CO2 capacity and selective adsorption over other small gases like N2, CH4, and O2, researchers have found success using strategies such as incorporation of coordinatively-unsaturated open metal sites (OMSs),17, 31-35 use of organic linkers featuring basic nitrogen-containing organic groups,36-41 and the introduction of strongly polarizing functional groups onto the linkers.42-45 Although these methods often increase the MOF-CO2 binding strength, they are associated with additional challenges. Many MOFs, 3

ACS Paragon Plus Environment

Chemistry of Materials

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 58

including those containing OMSs, are typically very sensitive to the presence of water vapor, which restricts many practical applications due to the ubiquity of humid environments.46-50 For example, MOF-74 incorporates OMSs and has exhibited a remarkable ability for CO2 capture; however, the M-MOF-74 series of MOFs (M = Zn, Co, Ni and Mg) loses most of its affinity for CO2 after exposure to humid environments due to preferential coordination of water molecules at guest binding sites.51-54 While the introduction of basic nitrogen-containing organic groups or other polarizing functional groups can increase CO2 adsorption capabilities in MOFs,31,

36-37, 55

these

procedures significantly increase the financial and labor costs of MOF synthesis, and reduce the commercial competitiveness of these materials compared to other porous media.56-57 Recently, several MOFs based on the semi-rigid, V-shaped H2SDB ligand (4,4̍sulfonyldibenzoic acid, Figure S1) have been reported.58-60 These systems have demonstrated high CO2/N2 selectivities even in humid environments,61-62 which is promising for practical applications. The four known H2SBD-based MOFs (CaSDB,58-59, 61 CdSDB,62 ZnSDB,63 and PbSDB64) all lack the OMSs or polarizing functional groups thought necessary for selective CO2 adsorption, indicating that a unique CO2 adsorption mechanism is present. The CO2 adsorption location and mechanism in any given MOF is strongly dependent on topological details and may also be influenced by the incorporated metal atoms, hence, detailed long- and short-range structural characterization is necessary to understand the interaction between CO2 and the MOF binding sites. An in-depth study of CO2 adsorption in these SDB-based MOFs is essential to understanding the exact guest locations, occupancies, dynamics, and binding strengths, in addition to the influences of the metal center and MOF topology on CO2 adsorption. The CO2 binding sites and adsorption mechanisms within PbSDB and CdSDB (Figure S2) are currently unknown. PbSDB was the first reported porous SDB-based MOF64 and is of special interest since a majority of Pb-based 4

ACS Paragon Plus Environment

Page 5 of 58

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

Chemistry of Materials

MOFs are non-porous.65 Although it is known that PbSDB has no uptake of N2 at 1 atm of pressure, CO2 adsorption has not been studied, and this MOF may also exhibit selective adsorption of CO2 over N2. CdSDB was recently reported62 and was shown to possess a high CO2/N2 selectivity, along with the ability to adsorb CO2 in conditions of 30 % humidity. Both PbSDB and CdSDB incorporate the V-shaped SDB ligand and feature one-dimensional channels, however, the metal center and coordination number has a profound influence on MOF topology: PbSDB features 7-coordinate Pb and linear channels, while the 6-coordinate Cd atoms in CdSDB and corresponding changes in topology impart a sinusoidal shape to the MOF channels (Figure S2). Since these MOFs share the same ligand but are based on different metals and display unique topologies, this is an excellent opportunity to examine the specific structural factors influencing CO2 adsorption in this system, and also to elucidate, compare, and contrast the locations and dynamic behavior of CO2 guests. Obtaining a comprehensive understanding of MOFs and their guests through a single characterization technique is elusive, if not impossible. An alternative approach is to perform X-ray diffraction (XRD) experiments and solid-state NMR (SSNMR) spectroscopy, using both of these powerful complementary solid characterization methods to investigate systems in great detail.66-67 Single crystal XRD (SCXRD) is a very useful technique for determining the long-range periodic crystal structure and differentiating between individual forms or phases of MOFs.68 In ideal conditions, SCXRD can determine the crystal structure and pinpoint the exact locations and site occupancies of guest molecules in MOFs. Unfortunately, conditions are not always ideal: SCXRD studies of guest-loaded MOFs are littered with challenges, including the growth of a high quality single crystal, preservation of crystal quality through the relatively harsh MOF activation process (i.e., solvent exchange, heating, vacuum), and loading sufficient amounts of guest molecules.69-73 In 5

ACS Paragon Plus Environment

Chemistry of Materials

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

addition, SCXRD only provides a time-averaged crystal structure; guest motion often leads to crystallographic disorder and uncertain site occupancies, limiting both the utility of the information and the depth of conclusions that can be obtained.68, 73 To minimize the problems associated with guest dynamics, SCXRD is commonly performed at low experimental temperatures, however, this restricts the amount of motional information that can be extracted (i.e., static guests at low temperatures necessarily yield no motional information). SSNMR can provide detailed information on guest dynamics and is a sensitive probe of short-range structure and crystallinity, serving as an excellent complement to SCXRD.66-67 SSNMR has been shown to yield key information regarding the framework structure,74-76 guest dynamics,77-78 and host–guest interactions79-80 in many MOFs. 13C and 1H SSNMR can be used to confirm the identity of the organic linkers and the presence of key functional groups upon modification of the linkers, and can monitor the adsorption of guest species.80-83 The metal ions within MOFs play a crucial role in structure-property relationships and are often NMR-active; accordingly, SSNMR of the metal centers provides rich information regarding the local geometry and coordination of the metal center, and can intimately probe metal-guest interactions.74, 82, 84-85 In situ variable-temperature (VT) SSNMR experiments can identify the specific types, rates, and routes of guest motion,78, 86-89 especially when there is detailed a priori knowledge of the MOF structure available from SCXRD experiments. The elegant combination of SSNMR and SCXRD is critical for investigating CO2 adsorption mechanisms, relating CO2 storage capacity to dynamics, discovering structural features conducive to CO2 adsorption, and ultimately designing better MOF-based CO2 absorbents. Herein, we describe a detailed SSNMR and SCXRD study of CO2 adsorption in PbSDB64 and CdSDB.62 The specific CO2 adsorption sites and the guest CO2 molecules themselves were 6

ACS Paragon Plus Environment

Page 6 of 58

Page 7 of 58

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

Chemistry of Materials

accurately located using SCXRD methods. In situ VT

13

C SSNMR techniques have been

successfully employed to reveal detailed information regarding the dynamics of CO2 guests in both of these SDB-based MOFs.

207

Pb and

111

Cd VT SSNMR techniques targeting the metal centers

reveal clear difference in CO2 guest locations and metal-guest interactions, which are connected to the unique Cd/PbSDB MOF structures and topologies. This study demonstrates that small alterations in MOF topology between two closely related systems can profoundly affect guest adsorption, and the knowledge of guest CO2 motion gained in this work has clear practical implications for adapting today and tomorrow’s MOFs for enhanced CO2 adsorption and storage.

Experimental Sample preparation. PbSDB was prepared according to the reported procedure,64 which is briefly summarized below. A reagent mixture of 0.165 g Pb(NO3)2 (BDH Chemicals, 99.5%), 0.306 g H2SDB (Sigma-Aldrich, 97%), and 0.123 g LiNO3 (BDH, 99.5%) was prepared in 10 mL of a 1:1 dimethylformamide (DMF, Caledon Laboratories, 99.8%) and methanol (MeOH, Fischer Chemical, 99.9%) solvent blend, and the resulting reagent mixture was placed into a Teflon chamber within a Teflon-lined stainless steel autoclave. The autoclave was then sealed, put into an oven, and heated at a temperature of 160 °C for 2 days. After cooling the autoclave to room temperature and allowing it to sit sealed without agitation for 5 days, colorless needle-shaped crystals were obtained, which were separated by vacuum filtration and washed with MeOH three times. CdSDB was prepared following the previously reported procedure62 with some modifications. A mixture of 0.109 g Cd(NO3)2 • 4H2O (Strem Chemicals, 98%) and 0.104 g H2SDB (Sigma-Aldrich, 97%) was dissolved in a mixture of 7 mL ethanol (EtOH, Fischer Chemical, 95 %) and 3 mL of deionized H2O. Following this, the reagent mixture was placed in a Teflon chamber 7

ACS Paragon Plus Environment

Chemistry of Materials

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 58

within a Teflon-lined stainless steel autoclave, which was then sealed and heated in a 180 °C oven for 2 days. After cooling the autoclave to room temperature and letting it rest without agitation for 48 hours, the product was obtained as colorless rectangular crystals, which were isolated by vacuum filtration and washed with EtOH three times. Calculated and experimental powder XRD patterns for both PbSDB and CdSDB are provided in the Supporting Information. Activation and CO2 loading of MOFs. To load Pb/CdSDB with CO2 and prepare these materials for SSNMR experiments, the following procedure was employed. The MOFs were first packed into the bottom of an L-shaped glass tube attached to a vacuum line, and a small amount of glass wool was used to secure the MOF sample in place. The glass tube was then heated at 180 °C / 453 K under dynamic vacuum (≤ 1 mbar) for ca. 12 hours to activate the MOF (i.e., purge all solvent and excess linker from the MOF channels). A known amount of

13

C-labeled

13

CO2

(Sigma-Aldrich, 99% 13C isotope enriched, referred to herein as simply CO2) was then introduced into the same empty vacuum line. CO2 was condensed and trapped in the MOF sample by immersing the glass tube in liquid nitrogen (77 K) and then flame-sealing the L-shaped tube off from the vacuum line. In the case of activated PbSDB loaded with 5% DMF, a set quantity of vaporized DMF was introduced to the vacuum line before flame-sealing the tube from the vacuum line. More details about the guest loading of all samples in this study are listed in Table S1. SSNMR measurements. All SSNMR experiments were performed using a Varian Infinity Plus wide-bore NMR spectrometer, equipped with an Oxford 9.4 T wide-bore magnet (ν0(13C) = 100.5 MHz, ν0(207Pb) = 83.7 MHz, ν0(111Cd) = 84.8 MHz) and a 5 mm HX static Varian/Chemagnetics probe unless otherwise noted. 13C NMR spectra were referenced to the high-frequency resonance of ethanol at 58.05 ppm.90 Static

13

C experiments were performed using the DEPTH-echo pulse 8

ACS Paragon Plus Environment

Page 9 of 58

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

Chemistry of Materials

sequence for enhanced background signal suppression,89, 91 employing a 90° pulse width of 3.1 µs and a 180° pulse width of 6.2 µs. The recycle delays and number of scans required for 13C SSNMR acquisition of both MOFs at variable temperatures are included in Supporting Information Table S2. Static

207

Pb NMR spectra were acquired using the WURST-CPMG pulse sequence92-97 with a 1H

decoupling field of 40 kHz, and were referenced to a saturated solution of Pb(NO3)2 at room temperature (δ(207Pb) = -2970 ppm).98 The calibrated 207Pb recycle delay employed was 10 s, while the spikelet separation in the frequency domain was 2500 Hz, and the spectral window was 1000 kHz. The static

111

Cd spectra were acquired using a variable-amplitude 1H-111Cd cross-polarized

(CP) echo pulse sequence, and referenced to 0.1 M Cd(ClO4)2 (δ(111Cd) = 0 ppm).99-100 The optimized recycle delay was 3 s, 1H 90º pulse was 3 µs, 1

111

Cd 180º pulse was 5.25 µs, and the

H-111Cd CP mixing time was 10 ms. 1H-111Cd CP-echo experiments also employed a spectral width

of 300 kHz, a Hartmann-Hahn matching field of 73 kHz, and a 1H decoupling field of 63 kHz. The 111

Cd isotope was chosen for study rather than

113

Cd due to rf interference from nearby radio

stations. 1

H-13C and 1H-111Cd CP magic angle spinning (CP/MAS) SSNMR experiments were performed

using a 4 mm HXY Varian/Chemagnetics probe and a spectral width of 100 kHz. 1H-13C CP/MAS experiments were performed at a spinning frequency of 14 kHz, using a contact time of 5 ms and a recycle delay of 3 s. 1H-13C CP experiments on PbSDB and CdSDB involved the acquisition of 298 and 1684 scans, respectively. 1H-111Cd CP/MAS experiments were carried out at a spinning frequency of 10 kHz using a contact time of 4 ms and a recycle delay of 1 s. 1204 scans were acquired on the as-made CdSDB sample, while 1072 scans were acquired for activated CdSDB. NMR chemical shift (CS) parameters. The 13C, 111Cd, and 207Pb SSNMR spectra in this study are dominated by the CS interaction, which describes the local magnetic environment about the 9

ACS Paragon Plus Environment

Chemistry of Materials

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

nucleus of interest. The CS interaction can be modeled by a second-rank tensor with three principal components δ11, δ22, and δ33, ordered such that δ11 ≥ δ22 ≥ δ33. We have employed the Herzfeld-Berger convention in this work, which uses three distinct parameters to describe the CS tensor. The isotropic chemical shift (δiso) is simply the average of the tensor components: δiso = 1 / 3 (δ11 + δ22 + δ33). The span (Ω) refers to the breadth of the powder pattern and is used to describe the chemical shift anisotropy (CSA), where Ω = δ11 - δ33. The axial symmetry of the CS tensor is modeled using the skew value (κ), defined as κ = 3(δ22 - δiso)/Ω; κ ranges from -1 to +1, with either limit corresponding to an axially symmetric CS tensor. SSNMR spectral simulations. The WSolids software package101 was used to analytically simulate all static 13C SSNMR spectra and extract the observed, or apparent, 13C NMR parameters of CO2 guests adsorbed within the MOFs. All errors in NMR parameters were estimated by bidirectional variation of the parameter of interest in both directions from the best-fit value. The EXPRESS software102 was used to simulate the effects of dynamic motion on 13C SSNMR spectra, assuming a linear geometry for the CO2 molecules along with the CS tensor parameters δ11 = δ22 = 245 ppm and δ33 = -90 ppm,103 and tiling across 4096 powder increments using the ZCW powder averaging procedure. Single crystal XRD. To begin, a small amount of single crystals were packed into NMR tubes, and activated using the same vacuum line and procedure as employed for activation prior to NMR experiments (vide supra). For MOF samples loaded to saturation with CO2, the glass tube containing the activated sample was exposed to CO2 while immersed in liquid nitrogen for ca. 15 minutes to ensure full saturation; the glass tube containing the CO2-saturated MOF was then flame-sealed from the vacuum line and kept at room temperature (Table S1). After one week at rest, the glass tubes were broken, and the CO2-saturated samples were immediately coated with paratone 10

ACS Paragon Plus Environment

Page 10 of 58

Page 11 of 58

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

Chemistry of Materials

oil in order to retain all CO2 within the MOF. An optical microscope was used to select high-quality single-crystals for structural analysis. For CdSDB samples loaded but not saturated with CO2, the glass tube containing activated CdSDB was only exposed to CO2 at liquid nitrogen temperature for 30 s before being flame-sealed from the vacuum line and stored at room temperature for 3 days (Table S1). All SCXRD measurements were made on a Bruker Kappa Axis Apex2 diffractometer at a temperature of 110 K. The frame integration was performed using SAINT.104 The resulting raw data was scaled and absorption corrected using a multi-scan averaging of symmetry equivalent data with SADABS.105 Structures were solved by using a dual space methodology incorporated in the SHELXT program.106 All non-hydrogen framework atoms were obtained from the initial structural solution. The hydrogen atoms were introduced at idealized positions and were allowed to ride on the parent atom. The atomic positions for the CO2 molecules were obtained from a difference Fourier map. The structural model was fit to the data using a full matrix least-squares method based on F2. The calculated structure factors included corrections for anomalous dispersion from the usual tabulation. The structure was refined using the SHELXL-2014 program from the SHELX suite of crystallographic software.107 A detailed summary of the crystallographic data of CO2-loaded Pb/CdSDB is given in Table S3 and Appendix A of the Supporting Information.

Results and Discussion In situ 13C SSNMR of PbSDB saturated with 13CO2. To understand the dynamics of saturated CO2 adsorbed within PbSDB, in situ 13C VT SSNMR experiments at temperatures ranging from 413 to 183 K were performed; the resulting experimental and analytically simulated spectra are illustrated in Figure 1. Since

13

C is only ca. 0.96 % naturally abundant,108 and given that 99 % 11

ACS Paragon Plus Environment

Chemistry of Materials

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

isotopically labeled

13

CO2 gas was used, all of the observed

Page 12 of 58

13

C resonances originate solely from

13

CO2 molecules rather than the framework.

Figure 1. The experimental static 13C VT SSNMR spectra of PbSDB loaded with saturated 13CO2 are shown at left in (a), and the corresponding spectral simulations to extract apparent NMR parameters are shown at right in (b). Since the MOF is loaded to saturation, there is excess non-adsorbed CO2, which corresponds to the sharp central resonance located at ca. 124 ppm. Note the disappearance of the central resonance corresponding to free CO2 at temperatures below 293 K.

Experimental

13

C SSNMR spectra obtained at temperatures from 413 to 293 K feature two

distinct resonances, (Figure 1(a)), as confirmed by analytical simulations (Figure 1(b)). The sharp, intense central resonance located at ca. 124 ppm originates from gaseous, mobile CO2, while the underlying

13

C powder pattern exhibits the same isotropic chemical shift (δiso) of 124 ppm but is

broadened by a considerable amount of chemical shift anisotropy (CSA) due to decreased CO2 mobility. The broad, temperature-dependent powder pattern undoubtedly corresponds to adsorbed within the MOF; similarly broad

13

CO2

13

C SSNMR powder patterns that change with

temperature have been observed for CO2 adsorbed in other MOFs.78, 86-87, 109 At temperatures below 293 K, the sharp

13

C resonance corresponding to gaseous CO2 is no

12

ACS Paragon Plus Environment

Page 13 of 58

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

Chemistry of Materials

longer visible, while the broad underlying powder pattern is clearly evident, implying that a majority of CO2 is adsorbed within the MOF. The broad powder pattern increases in width as temperature decreases, reflecting a further reduction in adsorbed CO2 mobility, and confirming that a significant adsorptive interaction exists between CO2 and PbSDB. At 183 K, the broad and well-defined single powder pattern is the sole spectral feature, indicating that only one CO2 adsorption site exists in PbSDB under these conditions and that no free CO2 remains. This finding is particularly interesting as there are no OMSs or polarizing functional groups in PbSDB, hinting that some other MOF feature gives rise to this adsorptive interaction. The span (Ω) value (i.e., powder pattern breadth) of adsorbed CO2 in PbSDB increases from 101(2) ppm at 413 K to 125(3) ppm at 183 K (Table 1); these span values are non-zero but less than half that of static CO2 (335 ppm).103 The moderate span exhibited by CO2 in PbSDB indicates that guest molecules are not static, nor completely mobile; the apparent

13

C span value and CSA is

significantly reduced by systematic CO2 motion about the adsorption site in PbSDB. As temperature decreases, the

13

C skew value (κ) decreases, suggesting that CO2 dynamics and perhaps the

MOF-CO2 interaction strength are altered. To gain a deeper understanding of CO2 motion in PbSDB, the EXPRESS software package102 was used to simulate experimental 13C SSNMR spectra and determine the types of dynamic motion present, along with their rates and associated motional angles. The simulations fit experimental data well (Figure 2(a)) and reveal that CO2 undergoes two types of distinct motion in PbSDB: a localized rotation upon the adsorption site modeled by a threefold (C3) rotation or “wobbling,” as well as a twofold (C2) non-localized hopping between symmetry-equivalent adsorption sites (Figure 2(b)); CO2 has been found to exhibit similar types of motion in other MOFs.78, 86-87, 109 We define the rotational angle between CO2 and the C3 wobbling axis as α, while the equivalent rotational angle 13

ACS Paragon Plus Environment

Chemistry of Materials

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

for C2 hopping about the rotational axis is denoted by β. It should be noted that the threefold wobbling is a model for a local rotation of CO2 through a continuum of orientations on the adsorption site defined by α, and all wobbling rotational symmetry ≥ C3 gives rise to the same powder pattern. The motional angles at each temperature are tabulated in Table 2. The motional rate of CO2 is in the fast motion regime (i.e., ≥ 107 Hz) at all temperatures. As the temperature increases from 183 to 383 K, the value of α increases slightly from 35(1) º to 39(1) º, indicating that CO2 wobbles through a larger angle and samples a larger volume of space when more thermal energy is available to compete with the adsorptive interaction. In contrast, the twofold hopping angle β decreases from 29(1) º at 183 K to 19(1) º at 383 K, which is likely due to a change in the orientation of CO2 molecules with respect to the C2 rotational axis as the experimental temperature is altered; a similar reduction in hopping angles with increased temperature was observed for guests in the MIL-53109 and MOF-74 frameworks.78, 88

Figure 2. In (a), simulated 13C SSNMR spectra incorporating both the wobbling and hopping motions are shown above experimental spectra, along with the threefold wobbling angles (α) and twofold hopping angles (β) determined at each temperature; both motions occur at a rate ≥ 107 Hz. 14

ACS Paragon Plus Environment

Page 14 of 58

Page 15 of 58

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

Chemistry of Materials

The * symbol denotes the resonance arising from free, unbound CO2. In (b), the two motion model of CO2 dynamics in PbSDB is depicted, which incorporates both local wobbling through an angle of α, and non-localized twofold hopping through an equivalent angle β between two symmetry-equivalent sites. Table 1. The apparent, or observed, 13C NMR parameters of adsorbed 13CO2 in SDB-based MOFs at different temperatures. MOF Adsorption site Temperature (K) δiso (ppm)a Ω (ppm) κ PbSDB 1 413 124(2) 101(2) 0.73(2) 383 124(2) 103(2) 0.70(3) 353 124(2) 105(2) 0.65(2) 323 124(2) 110(2) 0.60(2) 293 124(2) 115(3) 0.55(2) 243 124(2) 115(2) 0.45(2) 213 124(2) 118(2) 0.35(2) 183 124(2) 125(3) 0.30(3) CdSDB 1 413 124(1) 90(2) 0.18(9) 383 124(1) 101(2) 0.22(5) 353 124(1) 103(2) 0.26(3) 323 124(1) 106(2) 0.29(2) 293 124(1) 112(1) 0.33(1) 243 124(1) 114(1) 0.43(1) 213 124(1) 114(1) 0.43(1) 183 124(1) 116(1) 0.48(1) 2 413 110(4) 117(3) -1.00(7) 383 110(4) 123(3) -1.00(7) 353 110(3) 131(2) -1.00(6) 323 110(2) 138(2) -1.00(3) 293 108(2) 143(2) -1.00(2) 243 105(1) 156(2) -1.00(2) 213 101(1) 167(2) -0.95(2) 183 90(1) 175(2) -0.95(2) a This parameter was fixed at 124 ppm for CO2 located at adsorption site 1 in both systems, since this corresponds to our observed δiso for free CO2. In contrast, obtaining high-quality spectral simulations for adsorption site 2 in CdSDB required that δiso not be constrained.

15

ACS Paragon Plus Environment

Chemistry of Materials

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

Table 2. The CO2 motional data obtained from EXPRESS102 simulations of 13C VT SSNMR spectra. The rate of all motions is ≥ 107 Hz. MOF Temperature α (C3 Wobbling β (C2 Hopping (K) angle, °) angle, °) PbSDB 383 39(1) 19(1) 353 39(1) 22(1) 323 38(1) 23(1) 293 38(1) 25(1) 243 36(1) 28(1) 213 36(1) 29(1) 183 35(1) 29(1) a CdSDB 413 48(4) 33(3) 383 40(2) 32(2) 353 38(1) 31(2) 323 37(1) 31(2) 293 37(1) 29(1) 243 35(1) 29(1) 213 33(1) 29(1) 183 32(1) 28(1) a This is motional data for CO2 at adsorption site 1; simulations of experimental data for CO2 at adsorption site 2 did not yield reliable motional data (see discussion in text). SCXRD of PbSDB saturated with CO2.

13

C SSNMR spectra and corresponding simulations

have revealed the nature of CO2 dynamics in PbSDB, however, the exact location of the adsorption site remains unclear. Accordingly, SCXRD was used to obtain the crystal structure of CO2-saturated PbSDB at 110 K. The structure of activated PbSDB was used as a starting model64 and the guest CO2 atomic positions were subsequently obtained from a difference Fourier map. The long-range crystal lattice of CO2-saturated PbSDB is shown in Figure 3(a): all CO2 molecules are located at well-defined positions inside the one-dimensional channels that run along the crystallographic a axis, and none of the CO2 molecules exhibit positional disorder at 110 K. The CO2 geometry is linear (∠O-C-O = 179.96 º), and the normalized occupancy for the CO2 molecule was refined to a value of 0.45. Within each channel and unit cell of the PdSDB MOF, only one crystallographically independent CO2 molecule is observed, as denoted by the carbon atom C1A and corresponding CO2 molecule in Figure 3(b); this is in good agreement with the single unique CO2 species indicated 16

ACS Paragon Plus Environment

Page 16 of 58

Page 17 of 58

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

Chemistry of Materials

from the sole powder pattern in 13C SSNMR spectra. The CO2 carbon atom is located in the area bordered by the V-shaped SDB linker, adsorbed at an equivalent distance (3.550 Å) from both SDB phenyl rings (Figure 3(b)). A study of the related CaSDB MOF suggested that the V-shaped “π-pocket” geometry formed by the SDB linkers (Figure S3) is associated with CO2 uptake,61 since it facilitates attractive interactions between the molecular quadrupole moment of CO2 and the delocalized π aromatic system of the phenyl rings.110 The CO2 adsorption sites and binding mechanisms within other SDB-based MOFs have not yet been determined, however, PbSDB incorporates the same SDB linker as CaSDB and shares similar features, including the presence of a V-shaped π-pocket. Our SCXRD experiments clearly show that adsorbed CO2 resides in the middle of the SDB-based π-pocket structure within PbSDB. Although SCXRD gives a static picture of the time-averaged atomic positions, the ORTEP depiction of the CO2 adsorption site (Figure S4) reveals that the CO2 molecule is associated with large thermal ellipsoids, confirming that the localized CO2 wobbling motion still occurs at 110 K. The twofold hopping motion of CO2 can be explained by examining the positions of two neighboring symmetry-equivalent CO2 molecules along the longitudinal direction of an individual PbSDB channel (Figure 3(c,d)). The distance between the carbon atoms of these adjacent CO2 molecules is 3.144 Å, which is short enough to permit twofold hopping of CO2 molecules between both positions; yet because the two CO2 adsorption sites and molecules are symmetry-equivalent, only a single broad

13

C SSNMR powder is observed. The combined CO2 motional and positional

information from SSNMR and SCXRD experiments is illustrated in Figure 3(e). Based on the success of this experimental approach for elucidating the position and dynamics of CO2 adsorbed in PbSDB, a similar strategy was utilized to examine CO2-loaded CdSDB.

17

ACS Paragon Plus Environment

Chemistry of Materials

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 58

Figure 3. In (a), the extended lattice of the CO2-loaded PbSDB crystal structure is depicted, as viewed along the crystallographic a axis. The local structure about the CO2 “π-pocket” adsorption site is shown in (b), with equivalent distances shown from the carbon atom of CO2 to both nearby phenyl groups of the SDB linker. In (c) and (d), two adjacent symmetry-equivalent adsorbed CO2 molecules are shown from different perspectives. Lastly, in (e), the proposed two-motion model for CO2 dynamics in PbSDB is illustrated, with CO2 locally wobbling and non-locally hopping between adjacent equivalent adsorption sites. Carbon is colored grey, oxygen is red, lead is purple, sulfur is yellow, and the carbon atoms corresponding to CO2 guests are green. Note that the unit cell axis orientation is given in the top-left of each illustration, and is not consistent throughout (a) – (e).

In situ

13

C SSNMR and guest dynamics in CO2-saturated CdSDB. Much like PbSDB,

CdSDB also lacks the OMSs or polarizing functional groups thought to be key for gas adsorption. One would assume that the CO2 adsorption site in CdSDB should resemble the adsorption site in PbSDB and give rise to a single 13C SSNMR powder pattern; however, Figure 4(a) reveals that two distinct broad powder patterns corresponding to adsorbed CO2 are present in the spectra of CO2-loaded CdSDB, in addition to the sharp third

13

13

C VT SSNMR

C resonance belonging to mobile,

18

ACS Paragon Plus Environment

Page 19 of 58

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

Chemistry of Materials

non-adsorbed CO2. The analytical simulations in Figure 4(b) confirm that two CO2 adsorption sites are present in CdSDB, denoted as adsorption sites 1 and 2, which must differ in local geometry and/or CO2 binding affinity due to their distinct 13C NMR parameters (Table 1). The observed 13C NMR parameters of both powder patterns undergo clear changes as temperature is reduced, implying that CO2 guests adsorbed at both sites are undergoing motion on the NMR timescale. There are striking similarities in both visual appearance and apparent

13

C NMR parameters

between CO2 adsorbed at adsorption site 1 in CdSDB and that of CO2 adsorbed at the sole adsorption site in PbSDB (Table 1), strongly suggesting that these adsorption sites share a common location: the V-shaped π-pocket of the SDB ligand. In contrast, CO2 at adsorption site 2 in CdSDB corresponds to a unique powder pattern and different

13

C NMR parameters, implying that this

resonance arises from CO2 adsorbed at a different location in the MOF. The

13

C powder pattern

linked to adsorption site 2 is considerably broader than the powder pattern of adsorption site 1 (Ω = 143(2) versus 112(1) ppm at 293 K), implying that CO2 molecules at adsorption site 2 may experience reduced motion and a relatively stronger binding interaction with the MOF (Table 1, Figure S5). Adsorption site 2 is the more populated of the two at high experimental temperatures, but interestingly, more CO2 is located at adsorption site 1 at temperatures below 293 K (Figure S6). Since

13

C NMR parameters indicate reduced CO2 motion and a presumably stronger MOF-CO2

interaction at adsorption site 2, the increased CO2 population at adsorption site 1 at low temperatures suggests that site 1 must be significantly more accessible to CO2 guests. The low accessibility of site 2 may originate from a number of factors, such as disordered or distorted local geometry about the adsorption site.

19

ACS Paragon Plus Environment

Chemistry of Materials

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

Figure 4. The experimental static 13C VT SSNMR spectra of CO2-saturated CdSDB are shown in (a), accompanied by the analytically simulated spectra to extract observed NMR parameters in (b). Experimental temperatures are indicated at left, and the * label denotes the resonance corresponding to free gaseous CO2. 20

ACS Paragon Plus Environment

Page 20 of 58

Page 21 of 58

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

Chemistry of Materials

The 13C VT SSNMR spectra of CO2-saturated CdSDB were simulated to extract CO2 motional information. CO2 molecules at adsorption site 1 in CdSDB, which is presumably in the SDB π-pocket, exhibit the same type of wobbling and hopping motions as CO2 adsorbed at the π-pocket in PbSDB (Figure 5 and Table 2), as modeled by a fast (i.e., motional rate ≥ 107 Hz) threefold (C3) rotation and twofold (C2) hopping, respectively. At site 1 in CdSDB, wobbling angles (α) decrease from 48(4) ° at 413 K to 32(1) ° at 183 K, while α values are similar and likewise decrease from 40(1) ° to 36(1) ° in PbSDB, owing to the similar SDB π-pocket environments between these adsorption sites. The hopping angles (β) are similar between adsorption site 1 in CdSDB and in PbSDB, but the trends diverge: at site 1 in CdSDB, β decreases from 33(3) ° at 413 K to 28(1) ° at 183 K, while across this temperature range in PbSDB, β increases from 24(1) ° to 30(1) °. The most likely explanation is that temperature-induced changes in CO2 orientation occur in order to maximize attractive interactions with adsorption site 1, leading to a gradual alteration of hopping angles with temperature; the differences in MOF topology and local adsorption site geometry between PbSDB and CdSDB may also have some significant influence on CO2 hopping angles.

21

ACS Paragon Plus Environment

Chemistry of Materials

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

Figure 5. The simulated 13C SSNMR spectra associated with adsorption site 1 in CdSDB are shown as the purple spectra, along with motional simulations to obtain the extracted threefold wobbling angles (α) and twofold hopping angles (β) at each temperature depicted by the dashed red lines. All motions occur at a rate ≥ 107 Hz. Motional simulations could not be successfully performed on the 13 C SSNMR powder patterns corresponding to CO2 adsorbed at adsorption site 2, see discussion in text.

CO2 adsorbed at site 2 in CdSDB gives rise to a distinct 13C SSNMR powder pattern and NMR parameters that unfortunately could not be properly simulated to extract dynamic information. We were unable to obtain a satisfactory practical and meaningful model to describe CO2 dynamics at adsorption site 2 despite many attempts involving alternate motional simulations along with numerous different angles and rates of motion. Additional attempts to fit the powder pattern of adsorption site 2 using motional simulations based on the parameters of CO2 adsorbed at site 1 also repeatedly failed. Taken together, the sum of

13

C NMR parameters and decidedly unique CO2

dynamics strongly indicates that adsorption site 2 in CdSDB is fundamentally different in local 22

ACS Paragon Plus Environment

Page 22 of 58

Page 23 of 58

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

Chemistry of Materials

environment and adsorption affinity versus the π-pocket structure located at site 1 and in PbSDB. Furthermore, the

13

C powder pattern of CO2 at adsorption site 2 is broader than that of CO2 at

adsorption site 1, indicating that CO2 is bound more strongly and is less mobile at adsorption site 2. Unfortunately,

13

C SSNMR experiments alone cannot conclusively locate CO2, describe its

associated motions, and examine the local geometry about adsorption site 2 in CdSDB. In order to further investigate CO2 adsorption in PbSDB and CdSDB,

207

Pb and

111

Cd SSNMR experiments

were performed on various forms of these MOFs. 207

Pb and 111Cd SSNMR experiments on PbSDB and CdSDB. SSNMR is often employed to

study the local metal environment, metal-linker coordination, and metal-guest interactions in MOFs.74, 80, 84, 111-114

207

Pb and

111

Cd SSNMR experiments were carried out on PbSDB and CdSDB

MOFs under various conditions, with the resulting spectra and simulations shown in Figure 6, Figure S7, and Figure S8, and accompanying NMR parameters summarized in Tables 3 and 4.

23

ACS Paragon Plus Environment

Chemistry of Materials

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

Figure 6. The static 207Pb and 111Cd SSNMR spectra of PbSDB and CdSDB in various conditions are depicted. In (a), the 207Pb WURST-CPMG SSNMR spectra of PbSDB loaded with various guests are shown. “PbSDB with 0.5 CO2” means that the molar ratio of CO2 guests to Pb is 0.5. In (b), the experimental and simulated spectra of as-made PbSDB and activated PbSDB are illustrated. The static VT 1H-111Cd CP-echo SSNMR spectra of as-made, activated, and CO2-loaded CdSDB (CO2 : metal molar ratio of 0.2) are the broad spectra depicted in (c). The room temperature 1 H-111Cd CP/MAS spectra of as-made and activated CdSDB are also included in (c) as the relatively narrower lineshapes shown in the second and fourth traces from the top, respectively, with the δiso(111Cd) value labeled and spinning sidebands marked by an asterisk (*). A simulation of the static 1 H-111Cd room temperature SSNMR spectrum of CdSDB loaded with 0.2 CO2 is also included in (c), with simulations of individual powder patterns in purple (termed “CdL”) and green (known as CdR), along with the overall simulation in dashed red lines. See Table S1 for further details regarding guest loading. 24

ACS Paragon Plus Environment

Page 24 of 58

Page 25 of 58

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

Chemistry of Materials

Table 3. The 207Pb NMR parameters of various PbSDB samples as obtained from simulations of static WURST-CPMG experiments. Sample δiso (ppm) Ω (ppm) κ Resonance 1 -2057(13) 1158(45) 0.91(4) As made Resonance 2 -2420(30) 1123(130) -1.00(7) Activated -2067(15) 1155(35) 1.00(2) Loaded with 5% water -2067(15) 1178(30) 0.92(3) Loaded with 5% methanol -2067(17) 1178(40) 0.92(4) Loaded with 5% DMF -2067(20) 1178(40) 0.92(2) Loaded with 0.5 CO2 (293 K) -2077(15) 1170(35) 0.95(5) Loaded with 0.5 CO2 (183 K) -2057(15) 1155(35) 0.99(3)

Table 4. The 111Cd NMR parameters of activated and CO2-loaded CdSDB, as obtained from static VT 1H-111Cd CP-echo experiments. “CdL” and “CdR” refer to the “left” (high-frequency) and “right” (low-frequency) cadmium NMR powder patterns. Sample Temperature Resonance δiso Ω κ (ppm) (ppm) (°° K) As made 293 CdR -53(1) 242(7) -0.23(4) Activated 293 CdR -71(3) 233(8) -0.22(2) Activated 153 CdR -66(4) 252(5) -0.22(3) CdL -24(4) 199(14) 0.02(7) Loaded with 0.2 CO2 293 CdR -78(4) 243(9) -0.21(3) CdL -25(3) 199(12) 0.02(4) Loaded with 0.2 CO2 273 CdR -78(3) 228(10) -0.21(3) CdL -23(4) 199(10) 0.02(5) Loaded with 0.2 CO2 253 CdR -78(2) 217(8) -0.18(2) -23(4) 192(11) 0.04(4) CdL Loaded with 0.2 CO2 213 CdR -74(3) 216(9) -0.18(2) CdL -23(3) 192(13) 0.04(6) Loaded with 0.2 CO2 153 CdR -74(3) 215(7) -0.2(2) All

207

Pb SSNMR spectra were acquired under static conditions using the WURST-CPMG

pulse sequence, which yields spectra composed of a series of spikelets that trace out the overall manifold of the 207Pb powder pattern. The 207Pb SSNMR spectrum of as-made PbSDB (Figure 6(a)) is of an irregular shape, and simulations (Figure 6(b)) confirm the presence of two individual powder patterns of similar breadth but opposing κ values (Table 3). After activation, only the κ = +1 207

Pb powder pattern remains, indicative of a single unique Pb site and in agreement with the crystal 25

ACS Paragon Plus Environment

Chemistry of Materials

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

structure of PbSDB (Figure S3). Therefore, in the

207

Page 26 of 58

Pb SSNMR spectrum of as-made PbSDB, the

second powder pattern corresponding to κ = -1 must arise from Pb centers that are proximate to, or otherwise influenced by, residual solvent molecules from synthesis. The presence of residual solvent molecules in as-made PbSDB has a surprisingly profound impact on the axial symmetry of the 207Pb CS tensor, as evidenced by the change in κ from one form of axial symmetry (κ = +1, δ11 = δ22) to the other (κ = -1, δ22 = δ33). There is also a significant change of ca. 360 ppm in δiso. Given the apparent sensitivity of the

207

Pb CS tensor to the presence of guest molecules, PbSDB samples

loaded with a variety of solvents and CO2 were examined via

207

Pb SSNMR to further probe

host-guest interactions, and to investigate the origins of the second 207Pb powder pattern in as-made PbSDB. Since the solvent used for PbSDB synthesis was a 1:1 mixture of methanol and DMF, the impact of separately loading 5% wt of methanol, DMF, and water into activated PbSDB was examined via 207Pb SSNMR (Figure 6(a)). The water- and methanol-loaded samples of PbSDB each feature a single

207

Pb powder pattern which corresponds to nearly the same

207

Pb NMR parameters

as activated empty PbSDB (Figure S7, Table 3), confirming that neither solvent has an appreciable influence on the local environment of Pb. In contrast, DMF-loaded PbSDB yields a

207

Pb powder

pattern resembling that of as-made PbSDB; this spectrum features a clear increase in spectral intensity in the low-frequency region between ca. -2500 and -3000 ppm that is not particularly well-defined. The lack of a clearly resolved second powder pattern is likely due to short-range disorder about the solvent-proximate Pb centers, owing to the re-introduction of solvent to the activated MOF. From these

207

Pb SSNMR spectra, it is clear that DMF gives rise to the unique

second powder pattern in as-made PbSDB. The SSNMR data indicates that in PbSDB, MOF-DMF

26

ACS Paragon Plus Environment

Page 27 of 58

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

Chemistry of Materials

interactions significantly impact the local environment about Pb; however, DMF does not directly coordinate to the metal itself, since there are no open metal sites after activation. VT 207Pb SSNMR was also used to probe the impact of CO2 guests on the local environment of Pb and to investigate the possibility of any Pb-CO2 interactions in PbSDB. The

207

Pb SSNMR

spectrum of CO2-loaded PbSDB at 293 K is strikingly similar to the spectrum of empty activated PbSDB (Figure 6(a)), and simulations (Figure S7) indicate that both spectra correspond to approximately the same

207

Pb NMR parameters (Table 3). The clear similarities between the

207

Pb

NMR parameters of activated and CO2-loaded PbSDB imply that CO2 adsorbed in the SDB π-pocket has a minute or negligible influence on the local Pb environment. Pb does not directly interact with adsorbed CO2 and the local environment about Pb is preserved. No Pb-CO2 adsorptive interactions in these systems have otherwise been documented, in good agreement with our findings. To determine if CO2 dynamics have any impact on the Pb metal center in PbSDB, VT

207

Pb

SSNMR experiments were performed on CO2-loaded PbSDB at 183 K, yielding a powder pattern unchanged from room temperature and corresponding to approximately the same NMR parameters (Figure 6(a), Figure S7, Table 3). The differences in NMR parameters between temperatures are small, within experimental uncertainties, and likely due solely to the well-documented temperature dependence of the

207

Pb chemical shift;98 CO2 dynamics thus have no effect on the Pb metal center

in PbSDB. To see if the same conclusions held true when a different metal center was incorporated, 111

Cd SSNMR experiments were performed on CdSDB. The static 1H-111Cd CP-echo SSNMR spectra of as-made and activated CdSDB at room

temperature exhibit similar lineshapes (Figure 6(c)), and simulations suggest the presence of one 111

Cd powder pattern, in accordance with the single crystallographically unique cadmium site in the 27

ACS Paragon Plus Environment

Chemistry of Materials

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

CdSDB crystal structure.62 1H-111Cd CP/MAS spectra (Figure 6(c)) clearly confirm the presence of a single resonance at ca. δiso = -52 ppm in as-made and activated CdSDB. Static experiments reveal that the

111

Cd CS parameters Ω and κ are similar between as-made and activated CdSDB,

suggesting that Cd binding and connectivity is preserved (i.e., no Cd-linker bonds have been made or broken). This data indicates that the local Cd environment is only slightly perturbed by the removal of residual solvent from the MOF pores, likely via minor differences in linker-metal bond lengths and angles or linker orientation. It should be noted that there is some discrepancy in measured δiso(111Cd) values for activated CdSDB between 1H-113C static CP-echo (-71.3 ppm) and CP/MAS experiments (-52.4 ppm), which may be due to the influence of frictional heating from the spinning sample rotor (νrot = 10 kHz, 4 mm rotor). When the temperature is reduced to 153 K, static 1

H-111Cd CP-echo experiments on activated CdSDB reveal δiso has again shifted to -66(4) ppm,

supporting the notion of a temperature-dependent 111Cd chemical shift in this system. The introduction of CO2 to activated CdSDB at room temperature causes the appearance of a second

111

Cd powder pattern, implying that two distinct local Cd environments are now present

(Figure 6(c)). For clarity, these two powder patterns are denoted as CdL and CdR (left (high-frequency) and right (low-frequency), respectively, see Figure S8). CdR and its associated 111

Cd NMR parameters (δiso = -78(4) ppm, Ω = 243(9) ppm , κ = -0.21(3)) in CO2-loaded CdSDB at

room temperature are strikingly similar to the single powder pattern observed in activated CdSDB (δiso = -71(3) ppm , Ω = 233(8) ppm, κ = -0.22(2)), suggesting that these resonances correspond to nearly identical local Cd environments. In contrast, the CdL powder pattern is linked to unique 111Cd CS parameters (δiso = -24(4) ppm , Ω = 199(14) ppm, κ = 0.02(7)), and represents a second distinct Cd environment perturbed from that of activated CdSDB and associated with adsorption of CO2 guests. 28

ACS Paragon Plus Environment

Page 28 of 58

Page 29 of 58

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

Chemistry of Materials

VT static 1H-111Cd CP-echo SSNMR experiments were performed on CO2-loaded CdSDB at reduced temperatures to probe the influence of CO2 dynamics on the Cd metal centers. Although 111

Cd NMR parameters associated with both Cd resonances are invariant with reduced temperature

aside from a 10 % decrease in Ω(CdR), the relative intensity of CdL versus CdR climbs from ca. 0.2 at 293 K to ca. 1.2 at 153 K (Table 4, Figures S8 and S9). Since more CO2 is adsorbed within CdSDB at lower temperatures, the evolution in the CdL/CdR intensity ratio indicates that the population of these Cd environments are directly linked to the amount of adsorbed CO2 and permits spectral assignment: CdL corresponds to Cd centers proximate to adsorbed CO2 guests in occupied MOF channels, while CdR is linked to Cd in relatively empty channels that are distant from any CO2. The distinct

111

Cd NMR parameters associated with CdL suggest the presence of minor

guest-induced changes in metal-linker bond angles and/or linker orientations, which would have a measurable impact on the 111Cd CS tensor. Based on VT

13

C SSNMR experiments targeting adsorbed CO2 in CdSDB (vide supra), we

have hypothesized that CO2 adsorption site 1 in CdSDB is similar to the V-shaped π-pocket formed by the SDB linkers in PbSDB, while adsorption site 2 in CdSDB must have a unique geometry and engage in a different CO2 adsorptive interaction. The unchanged

111

Cd NMR parameters across

temperature variations in CO2-loaded CdSDB indicate that CO2 dynamics at sites 1 and 2 have very little or no influence on the Cd metal center, much like in the case of PbSDB. The impact of CO2 adsorption on 111Cd SSNMR spectra and associated NMR parameters implies adsorption site 2 may be located closer to Cd, but the question remains: assuming that the SDB-based π-pocket is already occupied by site 1, where exactly is adsorption site 2 located in CdSDB? Since the position of the CO2 adsorption sites in CdSDB are not readily available from our NMR data, SCXRD experiments were performed. 29

ACS Paragon Plus Environment

Chemistry of Materials

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 30 of 58

SCXRD of CO2-saturated CdSDB..The first sample examined was CdSDB loaded with saturated CO2 (Table S1) at a temperature of 110 K, which yielded the crystal structure pictured in Figure 7(a). All CO2 molecules are located inside the sinusoidal channels that propagate along the crystallographic c axis. There are two unique CO2 molecules in each pore, represented by their carbon atoms C1A and C1B (Figure 7(b)) that are located ca. 0.79 Å apart. There are two equivalent C1B sites and only one C1A site. The number of unique CO2 sites and their respective local environments are in agreement with our

13

C SSNMR data. The C1A site resides in the middle of the π-pocket structure formed by the

SDB linker, and is an equivalent 3.805 Å from both phenyl rings (Figure 7(c)). This crystallographic C1A site corresponds to the CO2 “adsorption site 1” resonance observed in

13

C

SSNMR spectra of CO2-loaded CdSDB (Figure 4). In this case, the twofold hopping between equivalent C1A sites extracted from SSNMR simulations likely describes CO2 jumping 5.75 Å down the longitudinal direction of the CdSDB channel (i.e., the c axis) between adjacent C1A adsorption sites, similar to the CO2 jumping in PbSDB. Based on the unique second powder pattern in

13

C SSNMR spectra and the influence of CO2

adsorption on 111Cd spectra, we postulated that the second CO2 adsorption site in CdSDB should be located relatively proximate to the Cd metal center. The measured distance between the nearest cadmium atom and C1B is 6.332 Å, versus a longer distance of 6.895 Å between C1A and the nearest cadmium center (Figure 7(d)). In a similar trend, the distance between C1B and the next nearest Cd is 6.379 Å, while the distance between the next nearest Cd and C1A is 7.007 Å. Therefore, the second CO2 adsorption site observed in 13C SSNMR spectra should correspond to the crystallographic site C1B, which is located relatively closer to Cd. Although 13C SSNMR experiments could not establish a motional model for CO2 at adsorption 30

ACS Paragon Plus Environment

Page 31 of 58

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

Chemistry of Materials

site 2 (vide supra), SCXRD data yields two important pieces of data. First, the measured distance between the two equivalent C1B carbon atoms in the same pore is 1.318 Å (Figure 7(d)), which is too close to permit CO2 hopping between adjacent C1B sites; any CO2 hopping at adsorption site 2 likely involves longitudinal jumping down the channel between C1B sites (i.e., along the c axis). Second, along with the short C1B-C1B distance of 1.318 Å, the distance between C1A and the equivalent C1B atoms is a short 0.79 Å; these short distances prohibit simultaneous occupation of all three adsorption sites in any given MOF channel cross-section.

Figure 7. The extended crystal structure of CdSDB loaded with saturated CO2, as seen along the c axis, is shown in (a). In (b), the local positions of adsorbed CO2 molecules in CdSDB are shown, highlighting the carbon atoms of CO2 that are associated with two unique adsorption sites C1A and C1B, where C1B has two symmetry-equivalent positions. The local environment about C1A is depicted in (c) along with its missing symmetry equivalent site located ca. 1.6 Å away, as indicated by a dashed red circle. The distance between C1A and both phenyl rings of the SDB linker are an equivalent 3.805 Å. In (d), the distances between C1B/C1A and adjacent cadmium atoms are shown, along with the short distance of 1.318 Å between the two symmetry-equivalent C1B adsorption sites. Note that the distance between C1B and both adjacent cadmium sites is significantly shorter 31

ACS Paragon Plus Environment

Chemistry of Materials

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

than those of C1A. Carbon is grey, oxygen is red, cadmium is blue, sulfur is yellow, and carbon atoms corresponding to adsorbed CO2 are green. Oxygen atoms on adsorbed CO2 are omitted for clarity and the purple phenyl rings in (b), (c), and (d) represent the other possible phenyl ring orientation in these disordered structures (see main text along with Figure S10).

An interesting question is why there is only a single C1A site indicated from SCXRD data, yet there are two SDB-based “π-pocket” structures capable of CO2 adsorption in CdSDB which should presumably give rise to two equivalent C1A sites (Figure 7(c)). We believe the absence of a second C1A adsorption site in SCXRD experiments is linked to our observation of disordered phenyl rings on the unique SDB linker that composes the “bottom” half of the channel perimeter in CdSDB (indicated by SCXRD experiments, see the purple carbon atoms in Figure 7(b,c,d)). It should be noted that the SDB linkers along the “top” half of the channel perimeter are not disordered. Disordered SDB phenyl rings were previously reported in CdSDB, but no detailed discussion was offered.62 The presence of disordered SDB linkers in CdSDB contrasts with the ordered phenyl rings in other SDB-based MOFs including PbSDB,59-60, 63, 115 however, disordered phenyls on linker species have been observed in other MOFs and coordination polymers.116-120 The specific activation process has been noted to influence the uptake of gases in a different SDB-based MOF,121 implying the possibility of a connection between linker orientation and gas uptake; however, our XRD data and prior reports62 confirm that disordered linkers are an intrinsic feature of CdSDB not connected to the activation process. The SDB phenyl groups contain carbon atoms C1, C2, C3, C4, C5, and C6 (all colored grey in Figure 7), and the disorder of the phenyl ring arises from a reorientation or partial flip around the C1-C4 axis, yielding alternate carbon positions C2', C3', C5', and C6' (all colored in purple in Figure 7). 1H-13C CP/MAS NMR experiments (Figure 8) confirm that there is static disorder of the phenyl rings in CdSDB: while PbSDB yields two

13

C resonances corresponding to the equivalent

32

ACS Paragon Plus Environment

Page 32 of 58

Page 33 of 58

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

Chemistry of Materials

C2-C6 and C3-C5 carbon atom pairs of the SDB phenyl ring, CdSDB yields four

13

C resonances

originating from the C2-C6, C3-C5, C2'-C6', and C3'-C5' pairs of SDB carbon atoms. The dihedral angle between the two phenyl planes is ca. 32.2º, and the normalized occupancy was refined to a value of 0.659(6) for the major conformer (depicted in grey). Owing to the disorder in phenyl orientations, there are four different combinations of phenyl ring orientations in the SDB ligand (Figure S10), however, only one combination can form a SDB-based π-pocket61 that affords simultaneous CO2 adsorptive interactions with both phenyl rings. Accordingly, CO2 adsorption and the CO2 population at this equivalent location to site 1 is greatly reduced; the resulting low density of CO2 molecules prohibit SCXRD detection of adsorbed guests at the lower π-pocket in the crystal structure (Figure 7(c), red dashed circle).

Figure 8. 1H-13C CP/MAS NMR experiments on activated (a) PbSDB and (b) CdSDB at a spinning frequency of 14 kHz. Note the extra 13C resonances in CdSDB located in the region from 137 – 145 ppm, which correspond to the other orientation of the static disordered phenyl rings. Further evidence for the static disorder is provided by SCXRD experiments, see Figure 7 and main text.

Although SCXRD data has shown that CO2 guests adsorbed in PbSDB adopt the typical linear 33

ACS Paragon Plus Environment

Chemistry of Materials

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

geometry, SCXRD indicates that adsorbed CO2 in CdSDB is bent in a nonlinear conformation. Bent CO2 molecules have been reported in a variety of different MOF systems and topologies,122-125 and the underlying details remain under debate.122 The bent CO2 configuration derived from SCXRD data is actually a time-averaged picture of CO2 structure, which can be significantly influenced by dynamics and the distribution of possible CO2 positions, and may not necessarily reflect the true CO2 geometry.122, 125 With these considerations in mind, we believe CO2 remains linear and do not place a high degree of confidence in the bent CO2 geometry observed in CdSDB; only the carbon positions are indicated in corresponding Figures. There are similar CO2 adsorption sites denoted C1A located at the π-pocket structures in both CdSDB and PbSDB, as shown by the green V shape in Figures 9(c, d), which exhibit very similar 13

C NMR powder patterns. In CdSDB there is a second CO2 adsorption site, C1B, that gives rise to

a unique

13

C powder pattern and is located at a different position in the channel (Figure 9(e)). A

more careful examination of the local geometry about C1B reveals that CO2 molecules located at this adsorption site may interact with a second type of π-pocket that is formed by two phenyl rings from the two separate SDB linkers attached to the cadmium metal centers. This second Cd-based π-pocket (Figure 9(e), blue V shape) is unique because the CO2 adsorption site is not equidistant from the phenyl rings: one C1B-phenyl distance is ca. 3.774 Å, and other C1B-phenyl distance is ca. 4.100 Å (Figure 9(e)). This second π-pocket does not exist in PbSDB since the phenyl group orientation about Pb makes such an arrangement impossible (Figure 9(f)). In the SDB-based π-pocket, the phenyl rings are in the correct orientation to bind CO2 molecules equidistant from both rings within a perpendicular plane. In contrast, the geometry about the Cd-based π-pocket is slightly distorted and the associated phenyl rings have deviated (i.e., tilted) away from an ideal orientation, resulting in an off-centered C1B adsorption site of unique local geometry that exhibits 34

ACS Paragon Plus Environment

Page 34 of 58

Page 35 of 58

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

Chemistry of Materials

different distances to each phenyl ring (Figure 9(e)); this is why CO2 adsorbed at the C1B site gives rise to a distinct second 13C powder pattern. It should be noted that the disordered phenyl rings of one SDB ligand in CdSDB (vide supra) may restrict CO2 adsorption at the C1B site in the Cd-based π–pocket to some extent due to unfavorable geometry; however, the successful location of both C1B sites via SCXRD methods clearly demonstrates that CO2 sufficiently populates both sites, and that phenyl ring disorder only has a minor impact on adsorption at the Cd-based π–pocket in CdSDB.

35

ACS Paragon Plus Environment

Chemistry of Materials

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

Figure 9. The extended crystal structure of CdSDB and PbSDB loaded with saturated CO2 are shown in (a) and (b), respectively. The local structure of CO2 molecules adsorbed at site 1 in CdSDB and at the only adsorption site within PbSDB are shown in (c) and (d). The carbon atoms of 36

ACS Paragon Plus Environment

Page 36 of 58

Page 37 of 58

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

Chemistry of Materials

CO2 adsorbed in both MOFs are located at equivalent distances from both nearby phenyl groups: for C1A in CdSDB, the distances are 3.805 Å, for C1A in PbSDB, the distances are 3.550 Å. The green “V” shape drawn above the phenyl rings of the SDB linker in (c), (d), and (f) highlights the V-shaped “π-pocket” associated with CO2 adsorption in these systems. The carbon atoms of CO2 found at the two symmetry-equivalent locations associated with adsorption site 2 in CdSDB (denoted C1B) are shown in (e), along with their different distances from the phenyl rings. The arrangement of phenyl groups about the cadmium centers gives rise to another V-shaped π-pocket, as indicated by the blue V shape in (e). In (f), the only π-pocket in PbSDB is pictured; the orientation of the bottom two SDB phenyl rings prevents formation of a second π-pocket structure. The carbon atoms associated with both unique CO2 adsorption sites in CdSDB are shown in (g), and the sole CO2 adsorption site along with its equivalent neighboring CO2 adsorption site are shown in (h). In all instances carbon is colored grey, oxygen is red, cadmium is blue, lead is purple, sulfur is yellow, and carbon atoms corresponding to adsorbed CO2 are green. Oxygen atoms on adsorbed CO2 in CdSDB are omitted for clarification, and the purple phenyl rings in (c), (e), and (g) represent another possible orientation of these disordered phenyl rings (see Figure S10).

The nature and coordination of the metal center has a profound influence on linker orientation and CO2 adsorption sites in these two SDB-based MOFs. Incorporation of the 7-coordinate Pb metal results in a single unique SDB-based CO2 adsorption site, as the framework topology forces the other unique SDB ligand to adopt a phenyl ring orientation that prevents π–pocket formation at this second SDB linker and at Pb. In contrast, selection of the 6-coordinate Cd center leads to two unique CO2 adsorption sites, since the MOF topology permits the appropriate SDB phenyl ring orientation to form one unique type of π–pocket in the V-shaped space of the SDB linker and two symmetry-equivalent π–pockets in the V-shaped spaces proximate to Cd. Our combined SCXRD and SSNMR data illustrates the importance of choosing the specific metal and its coordination number for the design and application SDB-based MOFs, since this directly influences MOF topology as well as the number, nature, and location of CO2 adsorption sites.

37

ACS Paragon Plus Environment

Chemistry of Materials

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

Figure 10. The extended crystal lattice of CdSDB loaded with unsaturated CO2 is depicted in (a), as viewed along the c axis. The locations of the symmetry-equivalent carbon atoms associated with CO2 adsorbed in CdSDB are shown in (b). Oxygen atoms on adsorbed CO2 are omitted for clarification, and the purple phenyl rings in (b) represent the other possible orientation of the disordered phenyl rings in CdSDB (see Figure S10). Carbon is colored grey, oxygen is red, cadmium is blue, sulfur is yellow, and carbon atoms corresponding to adsorbed CO2 are green.

SCXRD of CO2-loaded (unsaturated) CdSDB.. With the locations and origins of adsorption sites in PbSDB and CdSDB now known, the relative adsorption strengths and preferential occupation of adsorption sites in CdSDB can now be explored. 13C SSNMR data has suggested that the C1B site is preferentially occupied over C1A at room temperature. To investigate which adsorption site is preferred by CO2 guests, low levels of CO2 were loaded in an unsaturated manner (Table S1) to a single crystal of CdSDB for SCXRD experiments, yielding the structure depicted in Figure 10(a). There are two equivalent C1B CO2 molecules located near their usual Cd-based π-pocket positions in the pore, while there are no C1A CO2 molecules to be found in either SDB-based π-pocket structure (Figure 10(b)). It should be noted that although the C1B in this instance remains positioned within the Cd-based π-pocket, C1B is also located slightly closer to the SDB-based π-pocket structure versus the C1B position within CO2-saturated CdSDB (refer to Appendix A and associated crystal structures). The normalized occupancy of C1B CO2 in unsaturated CdSDB was refined to a value of 0.69. These SCXRD experiments clearly show that a 38

ACS Paragon Plus Environment

Page 38 of 58

Page 39 of 58

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

Chemistry of Materials

majority of CO2 is adsorbed at the C1B adsorption site when CO2 loading levels are low, indicating that CO2 adsorption strength is likely higher at the C1B site in the Cd-based π–pocket, and binding should be relatively weaker at the SDB-based π-pocket associated with C1A. Of particular note is that the relatively stronger CO2 adsorption associated with the Cd-based π–pocket is linked to unequal CO2-phenyl distances, while the weaker adsorption at the SDB-based π–pocket involves equal CO2-phenyl distances; this raises the potential for further increasing the CO2 binding energy in SDB-based MOFs, and MOFs in general, by modification of specific phenyl ring orientations and distances. Much like the situation of saturated CO2 loaded in CdSDB, the geometry of unsaturated CO2 in CdSDB was also found to be nonlinear in SCXRD experiments. In addition to the previously discussed reasons (vide supra), another reason for the observation of bent CO2 in unsaturated CdSDB may be the low electron density of adsorbed CO2 arising from the relatively low unsaturated guest loading level. Finally, SCXRD data was thoroughly examined in order to understand the influence of adsorbed CO2 at adsorption site 2 (C1B) on the Cd local environment and 111Cd NMR parameters. A comparison of activated empty CdSDB and CO2-loaded CdSDB SCXRD data (Table S4) reveals that the unit cell parameters and local Cd coordination environment (including bond lengths and angles) are nearly the same in both systems, hinting some other structural change gives rise to the changes in

111

Cd NMR parameters. Since each Cd center is bound to three SDB linkers with

disordered phenyl rings (six connected SDB linkers in total, see Figure S3), it is possible that adsorbed CO2 at site 2 (C1B) may influence the two possible orientations of the nearby disordered phenyl groups associated with the Cd-based π–pocket. In activated CdSDB, SCXRD data indicates that the dihedral angle between the two possible orientations of phenyl rings (Figure S11, Table S5) are identical for both phenyl rings attached to 39

ACS Paragon Plus Environment

Chemistry of Materials

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 SDB linker. After loading CO2 into CdSDB at unsaturated levels, the dihedral angles between the disordered phenyl orientations for both phenyl groups in the same SDB linker now differ by ca. 0.3°, and are also altered from the values in empty activated CdSDB (Table S5), suggesting that CO2 guests have some influence on phenyl orientation. In CO2-saturated CdSDB, the disparity between the measured dihedral angles grows larger to ca. 0.6°. These findings help explain the two observed powder patterns in 1H-111Cd CP SSNMR spectra (Figure 6(c)): the presence of guest CO2 slightly modifies the individual static orientations of the disordered phenyl rings on the SDB ligand, which perturbs the

111

Cd NMR parameters and gives rise to the CdL resonance. When more CO2 is

adsorbed in CdSDB (i.e., at lower experimental temperatures), the phenyl ring orientations are modified to a greater degree, creating more CdL environments and causing a further increase in CdL powder pattern intensity. In this case, even small changes in dihedral angles between the static disordered phenyl orientations can be clearly detected from both SCXRD and SSNMR methods.

Conclusions A series of SSNMR and SCXRD experiments have been used to understand CO2 adsorption and dynamics in the PbSDB and CdSDB MOFs, which do not feature open metal sites or polarizing functional groups. Multinuclear VT SSNMR experiments revealed the number of CO2 adsorption sites in each MOF and yielded detailed guest motional data, as well as information on the MOF-CO2 adsorptive interaction and the influence of the metal center on CO2 adsorption. CO2 guests undergo a combination of a localized rotation or “wobbling” as well as a non-localized hopping between symmetry-equivalent SDB-based adsorption sites in both MOFs. SCXRD experiments were used to elucidate the exact locations of CO2 adsorption sites and their occupancies, along with the relationship between phenyl ring disorder and guest adsorption. CO2 40

ACS Paragon Plus Environment

Page 40 of 58

Page 41 of 58

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

Chemistry of Materials

adsorption occurs at the V-shaped π–pocket in both MOFs, which is located equidistant from both phenyl rings of the bent SDB linker. SSNMR and SCXRD experiments reveal that the second adsorption site in CdSDB is located in another π-pocket formed by phenyl rings about the 6-coordinate Cd metal; in contrast, the distinct connectivity between 7-coordinate Pb and SDB linkers in PbSDB prohibits the phenyl ring geometry necessary for the formation of an additional Pb-based π-pocket and a second CO2 adsorption site. The choice of metal center and coordination number permits controlled incorporation of one or multiple π-pocket structures in a single MOF, which has the potential to increase CO2 adsorption and permit fine-tuning of the guest adsorption strength within future rationally-designed MOFs. Both SSNMR and SCXRD experiments detected uncommon CO2-induced structural changes in CdSDB, which provide yet another avenue to further optimize CO2 adsorption in MOFs. This work is a vivid illustration of how diverse, detailed structural and motional insights are available when utilizing a combination of solid state NMR and SCXRD experiments; pairing these complementary characterization methods is an extremely effective avenue for studying guest adsorption in MOFs.

Supporting Information. The Supporting Information is available free of charge on the ACS Publications website at DOI: __________. Additional structural diagrams, experimental details, detailed single-crystal XRD information, multinuclear (13C,

111

Cd,

207

Pb) SSNMR spectra and

associated parameters, powder XRD patterns, optical microscope images, and ORTEP illustrations of single-crystal XRD structures.

Acknowledgements Y. H. thanks the Natural Science and Engineering Research Council (NSERC) of Canada for a 41

ACS Paragon Plus Environment

Chemistry of Materials

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

Discovery Grant and a Discovery Accelerator Supplements Award.

References 1. Solomon, S.; Plattner, G.-K.; Knutti, R.; Friedlingstein, P., Irreversible Climate Change Due to Carbon Dioxide Emissions. Proc. Natl. Acad. Sci. 2009, 106, 1704-1709. 2. Hansen, J.; Sato, M.; Ruedy, R.; Lo, K.; Lea, D. W.; Medina-Elizade, M., Global Temperature Change. Proc. Natl. Acad. Sci. 2006, 103, 14288-14293. 3. Thomas, C. D.; Cameron, A.; Green, R. E.; Bakkenes, M.; Beaumont, L. J.; Collingham, Y. C.; Erasmus, B. F.; De Siqueira, M. F.; Grainger, A.; Hannah, L., Extinction Risk from Climate Change. Nature 2004, 427, 145-148. 4. Patz, J. A.; Campbell-Lendrum, D.; Holloway, T.; Foley, J. A., Impact of Regional Climate Change on Human Health. Nature 2005, 438, 310-317. 5. Şekercioğlu, Ç. H.; Primack, R. B.; Wormworth, J., The Effects of Climate Change on Tropical Birds. Biol. Cons. 2012, 148, 1-18. 6. Duarte, H.; Tejedo, M.; Katzenberger, M.; Marangoni, F.; Baldo, D.; Beltrán, J. F.; Martí, D. A.; Richter‐Boix, A.; Gonzalez‐Voyer, A., Can Amphibians Take the Heat? Vulnerability to Climate Warming in Subtropical and Temperate Larval Amphibian Communities. Global Change Biol. 2012, 18, 412-421. 7. Urban, M. C., Accelerating Extinction Risk from Climate Change. Science 2015, 348, 571-573. 8. D'Alessandro, D. M.; Smit, B.; Long, J. R., Carbon Dioxide Capture: Prospects for New Materials. Angew. Chem. Int. Ed. 2010, 49, 6058-6082. 9. 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. 2011, 112, 42

ACS Paragon Plus Environment

Page 42 of 58

Page 43 of 58

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

Chemistry of Materials

724-781. 10. Furukawa, H.; Cordova, K. E.; O’Keeffe, M.; Yaghi, O. M., The Chemistry and Applications of Metal-Organic Frameworks. Science 2013, 341, 1230444(1)-1230444(12). 11. O’Keeffe, M.; Yaghi, O. M., Deconstructing the Crystal Structures of Metal–Organic Frameworks and Related Materials into Their Underlying Nets. Chem. Rev. 2011, 112, 675-702. 12. Gándara, F.; Furukawa, H.; Lee, S.; Yaghi, O. M., High Methane Storage Capacity in Aluminum Metal-Organic Frameworks. J. Am. Chem. Soc. 2014, 136, 5271-5274. 13. Tranchemontagne, D. J.; Park, K. S.; Furukawa, H.; Eckert, J.; Knobler, C. B.; Yaghi, O. M., Hydrogen Storage in New Metal-Organic Frameworks. J. Phys. Chem. C 2012, 116, 13143-13151. 14. Murray, L. J.; Dincă, M.; Long, J. R., Hydrogen Storage in Metal–Organic Frameworks. Chem. Soc. Rev. 2009, 38, 1294-1314. 15. Feng, D.; Wang, K.; Wei, Z.; Chen, Y.-P.; Simon, C. M.; Arvapally, R. K.; Martin, R. L.; Bosch, M.; Liu, T.-F.; Fordham, S., Kinetically Tuned Dimensional Augmentation as a Versatile Synthetic Route Towards Robust Metal-Organic Frameworks. Nature Commun. 2014, 5, 5723(1)-5723(8). 16. Das, M. C.; Guo, Q.; He, Y.; Kim, J.; Zhao, C.-G.; Hong, K.; Xiang, S.; Zhang, Z.; Thomas, K. M.; Krishna, R., Interplay of Metalloligand and Organic Ligand to Tune Micropores within Isostructural Mixed-Metal Organic Frameworks (M′MOFs) for Their Highly Selective Separation of Chiral and Achiral Small Molecules. J. Am. Chem. Soc. 2012, 134, 8703-8710. 17. Li, J.-R.; Sculley, J.; Zhou, H.-C., Metal-Organic Frameworks for Separations. Chem. Rev. 2011, 112, 869-932. 18. Herm, Z. R.; Wiers, B. M.; Mason, J. A.; van Baten, J. M.; Hudson, M. R.; Zajdel, P.; Brown, C. M.; Masciocchi, N.; Krishna, R.; Long, J. R., Separation of Hexane Isomers in a Metal-Organic Framework with Triangular Channels. Science 2013, 340, 960-964. 43

ACS Paragon Plus Environment

Chemistry of Materials

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

19. Na, K.; Choi, K. M.; Yaghi, O. M.; Somorjai, G. A., Metal Nanocrystals Embedded in Single Nanocrystals of MOFs Give Unusual Selectivity as Heterogeneous Catalysts. Nano Lett. 2014, 14, 5979-5983. 20. Dhakshinamoorthy, A.; Alvaro, M.; Garcia, H., Commercial Metal-Organic Frameworks as Heterogeneous Catalysts. Chem. Commun. 2012, 48, 11275-11288. 21. Chen, B.; Wang, L.; Xiao, Y.; Fronczek, F. R.; Xue, M.; Cui, Y.; Qian, G., A Luminescent Metal-Organic Framework with Lewis Basic Pyridyl Sites for the Sensing of Metal Ions. Angew. Chem. Int. Ed. 2009, 48, 500-503. 22. Kreno, L. E.; Leong, K.; Farha, O. K.; Allendorf, M.; Van Duyne, R. P.; Hupp, J. T., Metal-Organic Framework Materials as Chemical Sensors. Chem. Rev. 2011, 112, 1105-1125. 23. Hu, Z.; Deibert, B. J.; Li, J., Luminescent Metal-Organic Frameworks for Chemical Sensing and Explosive Detection. Chem. Soc. Rev. 2014, 43, 5815-5840. 24. Horcajada, P.; Gref, R.; Baati, T.; Allan, P. K.; Maurin, G.; Couvreur, P.; Férey, G.; Morris, R. E.; Serre, C., Metal-Organic Frameworks in Biomedicine. Chem. Rev. 2011, 112, 1232-1268. 25. Cunha, D.; Ben Yahia, M.; Hall, S.; Miller, S. R.; Chevreau, H.; Elkaïm, E.; Maurin, G.; Horcajada, P.; Serre, C., Rationale of Drug Encapsulation and Release from Biocompatible Porous Metal-Organic Frameworks. Chem. Mater. 2013, 25, 2767-2776. 26. Kundu, T.; Mitra, S.; Patra, P.; Goswami, A.; Díaz Díaz, D.; Banerjee, R., Mechanical Downsizing of a Gadolinium (III)‐Based Metal-Organic Framework for Anticancer Drug Delivery. Chem. Eur. J. 2014, 20, 10514-10518. 27. McDonald, T. M.; Lee, W. R.; Mason, J. A.; Wiers, B. M.; Hong, C. S.; Long, J. R., Capture of Carbon Dioxide from Air and Flue Gas in the Alkylamine-Appended Metal-Organic Framework mmen-Mg2 (dobpdc). J. Am. Chem. Soc. 2012, 134, 7056-7065. 44

ACS Paragon Plus Environment

Page 44 of 58

Page 45 of 58

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

Chemistry of Materials

28. Xiang, S.; He, Y.; Zhang, Z.; Wu, H.; Zhou, W.; Krishna, R.; Chen, B., Microporous Metal-Organic Framework with Potential for Carbon Dioxide Capture at Ambient Conditions. Nature Commun. 2012, 3, 954(1)-954(9). 29. Zhang, H.; Jin, M.; Liu, H.; Wang, J.; Kim, M. J.; Yang, D.; Xie, Z.; Liu, J.; Xia, Y., Facile Synthesis of Pd-Pt Alloy Nanocages and Their Enhanced Performance for Preferential Oxidation of CO in Excess Hydrogen. ACS Nano 2011, 5, 8212-8222. 30. Silva, P.; Vilela, S. M.; Tomé, J. P.; Paz, F. A. A., Multifunctional Metal-Organic Frameworks: From Academia to Industrial Applications. Chem. Soc. Rev. 2015, 44, 6774-6803. 31. Mason, J. A.; Sumida, K.; Herm, Z. R.; Krishna, R.; Long, J. R., Evaluating Metal-Organic Frameworks for Post-Combustion Carbon Dioxide Capture Via Temperature Swing Adsorption. Energy Environ. Sci. 2011, 4, 3030-3040. 32. Bae, Y.-S.; Farha, O. K.; Spokoyny, A. M.; Mirkin, C. A.; Hupp, J. T.; Snurr, R. Q., Carborane-Based Metal-Organic Frameworks as Highly Selective Sorbents for CO2 over Methane. Chem. Commun. 2008, 4135-4137. 33. Li, B.; Zhang, Z.; Li, Y.; Yao, K.; Zhu, Y.; Deng, Z.; Yang, F.; Zhou, X.; Li, G.; Wu, H., Enhanced Binding Affinity, Remarkable Selectivity, and High Capacity of CO2 by Dual Functionalization of a Rht‐Type Metal-Organic Framework. Angew. Chem. Int. Ed. 2012, 51, 1412-1415. 34. Xue, D.-X.; Cairns, A. J.; Belmabkhout, Y.; Wojtas, L.; Liu, Y.; Alkordi, M. H.; Eddaoudi, M., Tunable Rare-Earth Fcu-MOFs: A Platform for Systematic Enhancement of CO2 Adsorption Energetics and Uptake. J. Am. Chem. Soc. 2013, 135, 7660-7667. 35. Herm, Z. R.; Swisher, J. A.; Smit, B.; Krishna, R.; Long, J. R., Metal-Organic Frameworks as Adsorbents for Hydrogen Purification and Precombustion Carbon Dioxide Capture. J. Am. Chem. 45

ACS Paragon Plus Environment

Chemistry of Materials

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 46 of 58

Soc. 2011, 133, 5664-5667. 36. Demessence, A.; D’Alessandro, D. M.; Foo, M. L.; Long, J. R., Strong CO2 Binding in a Water-Stable, Triazolate-Bridged Metal−Organic Framework Functionalized with Ethylenediamine. J. Am. Chem. Soc. 2009, 131, 8784-8786. 37. Lu, W.; Sculley, J. P.; Yuan, D.; Krishna, R.; Wei, Z.; Zhou, H. C., Polyamine-Tethered Porous Polymer Networks for Carbon Dioxide Capture from Flue Gas. Angew. Chem. Int. Ed. 2012, 51, 7480-7484. 38. Couck, S.; Denayer, J. F.; Baron, G. V.; Rémy, T.; Gascon, J.; Kapteijn, F., An Amine-Functionalized MIL-53 Metal- Organic Framework with Large Separation Power for CO2 and CH4. J. Am. Chem. Soc. 2009, 131, 6326-6327. 39. Serra-Crespo, P.; Ramos-Fernandez, E. V.; Gascon, J.; Kapteijn, F., Synthesis and Characterization of an Amino Functionalized MIL-101 (Al): Separation and Catalytic Properties. Chem. Mater. 2011, 23, 2565-2572. 40. Zheng, B.; Bai, J.; Duan, J.; Wojtas, L.; Zaworotko, M. J., Enhanced CO2 Binding Affinity of a High-Uptake Rht-Type Metal-Organic Framework Decorated with Acylamide Groups. J. Am. Chem. Soc. 2010, 133, 748-751. 41. Rabbani,

M.

G.;

El-Kaderi,

H.

M.,

Synthesis

and

Characterization

of

Porous

Benzimidazole-Linked Polymers and Their Performance in Small Gas Storage and Selective Uptake. Chem. Mater. 2012, 24, 1511-1517. 42. Neofotistou, E.; D Malliakas, C.; N Trikalitis, P., Unprecedented Sulfone-Functionalized Metal-Organic Frameworks and Gas-Sorption Properties. Chem. Eur. J. 2009, 15, 4523-4527. 43. Tan, Y.-X.; Wang, F.; Kang, Y.; Zhang, J., Dynamic Microporous Indium (III)-4, 4′-Oxybis (Benzoate) Framework with High Selectivity for the Adsorption of CO2 over N2. Chem. Commun. 46

ACS Paragon Plus Environment

Page 47 of 58

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

Chemistry of Materials

2011, 47, 770-772. 44. Luo, F.; Wang, M.-S.; Luo, M.-B.; Sun, G.-M.; Song, Y.-M.; Li, P.-X.; Guo, G.-C., Functionalizing the Pore Wall of Chiral Porous Metal-Organic Frameworks by Distinct –H, –OH, – NH2, –NO2, –COOH Shutters Showing Selective Adsorption of CO2, Tunable Photoluminescence, and Direct White-Light Emission. Chem. Commun. 2012, 48, 5989-5991. 45. Wang, B.; Huang, H.; Lv, X.-L.; Xie, Y.; Li, M.; Li, J.-R., Tuning CO2 Selective Adsorption over N2 and CH4 in UiO-67 Analogues through Ligand Functionalization. Inorg. Chem. 2014, 53, 9254-9259. 46. Kaye, S. S.; Dailly, A.; Yaghi, O. M.; Long, J. R., Impact of Preparation and Handling on the Hydrogen Storage Properties of Zn4O(1, 4-Benzenedicarboxylate)3 (MOF-5). J. Am. Chem. Soc. 2007, 129, 14176-14177. 47. Canivet, J.; Fateeva, A.; Guo, Y.; Coasne, B.; Farrusseng, D., Water Adsorption in MOFs: Fundamentals and Applications. Chem. Soc. Rev. 2014, 43, 5594-5617. 48. Choi, S.; Watanabe, T.; Bae, T.-H.; Sholl, D. S.; Jones, C. W., Modification of the Mg/DOBDC MOF with Amines to Enhance CO2 Adsorption from Ultradilute Gases. J. Phys. Chem. Lett. 2012, 3, 1136-1141. 49. Schoenecker, P. M.; Carson, C. G.; Jasuja, H.; Flemming, C. J.; Walton, K. S., Effect of Water Adsorption on Retention of Structure and Surface Area of Metal–Organic Frameworks. Ind. Eng. Chem. Res. 2012, 51, 6513-6519. 50. Mottillo, C.; Friščić, T., Carbon Dioxide Sensitivity of Zeolitic Imidazolate Frameworks. Angew. Chem. Int. Ed. 2014, 53, 7471-7474. 51. Kizzie, A. C.; Wong-Foy, A. G.; Matzger, A. J., Effect of Humidity on the Performance of Microporous Coordination Polymers as Adsorbents for CO2 Capture. Langmuir 2011, 27, 47

ACS Paragon Plus Environment

Chemistry of Materials

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

6368-6373. 52. Liu, J.; Wang, Y.; Benin, A. I.; Jakubczak, P.; Willis, R. R.; LeVan, M. D., CO2/H2O Adsorption Equilibrium and Rates on Metal-Organic Frameworks: HKUST-1 and Ni/DOBDC. Langmuir 2010, 26, 14301-14307. 53. Kumar, A.; Madden, D. G.; Lusi, M.; Chen, K. J.; Daniels, E. A.; Curtin, T.; Perry, J. J.; Zaworotko, M. J., Direct Air Capture of CO2 by Physisorbent Materials. Angew. Chem. Int. Ed. 2015, 54, 14372-14377. 54. Woerner, W. R.; Plonka, A. M.; Chen, X.; Banerjee, D.; Thallapally, P. K.; Parise, J. B., Simultaneous in Situ X-Ray Diffraction and Calorimetric Studies as a Tool to Evaluate Gas Adsorption in Microporous Materials. J. Phys. Chem. C. 2015, 120, 360-369. 55. Dietzel, P. D.; Besikiotis, V.; Blom, R., Application of Metal–Organic Frameworks with Coordinatively Unsaturated Metal Sites in Storage and Separation of Methane and Carbon Dioxide. J. Mater. Chem. 2009, 19, 7362-7370. 56. Liu, J.; Thallapally, P. K.; McGrail, B. P.; Brown, D. R.; Liu, J., Progress in Adsorption-Based CO2 Capture by Metal–Organic Frameworks. Chem. Soc. Rev. 2012, 41, 2308-2322. 57. Zhang, Z.; Zhao, Y.; Gong, Q.; Li, Z.; Li, J., MOFs for CO2 Capture and Separation from Flue Gas Mixtures: The Effect of Multifunctional Sites on Their Adsorption Capacity and Selectivity. Chem. Commun. 2013, 49, 653-661. 58. Banerjee, D.; Zhang, Z.; Plonka, A. M.; Li, J.; Parise, J. B., A Calcium Coordination Framework Having Permanent Porosity and High CO2/N2 Selectivity. Cryst. Growth Des. 2012, 12, 2162-2165. 59. Yeh, C.-T.; Lin, W.-C.; Lo, S.-H.; Kao, C.-C.; Lin, C.-H.; Yang, C.-C., Microwave Synthesis and Gas Sorption of Calcium and Strontium Metal–Organic Frameworks with High Thermal 48

ACS Paragon Plus Environment

Page 48 of 58

Page 49 of 58

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

Chemistry of Materials

Stability. CrystEngComm 2012, 14, 1219-1222. 60. Senthil Raja, D.; Luo, J.-H.; Wu, C.-Y.; Cheng, Y.-J.; Yeh, C.-T.; Chen, Y.-T.; Lo, S.-H.; Lai, Y.-L.; Lin, C.-H., Solvothermal Synthesis, Structural Diversity, and Properties of Alkali Metal– Organic Frameworks Based on V-Shaped Ligand. Cryst. Growth Des. 2013, 13, 3785-3793. 61. Plonka, A. M.; Banerjee, D.; Woerner, W. R.; Zhang, Z.; Nijem, N.; Chabal, Y. J.; Li, J.; Parise, J. B., Mechanism of Carbon Dioxide Adsorption in a Highly Selective Coordination Network Supported by Direct Structural Evidence. Angew. Chem. Int. Ed. 2013, 52, 1692-1695. 62. Plonka, A. M.; Banerjee, D.; Woerner, W. R.; Zhang, Z.; Li, J.; Parise, J. B., Effect of Ligand Geometry on Selective Gas-Adsorption: The Case of a Microporous Cadmium Metal Organic Framework with a V-Shaped Linker. Chem. Commun. 2013, 49, 7055-7057. 63. Wu, C.-Y.; Raja, D. S.; Yang, C.-C.; Yeh, C.-T.; Chen, Y.-R.; Li, C.-Y.; Ko, B.-T.; Lin, C.-H., Evaluation of Structural Transformation in 2D Metal-Organic Frameworks Based on a 4, 4′-Sulfonyldibenzoate Linker: Microwave-Assisted Solvothermal Synthesis, Characterization and Applications. CrystEngComm 2014, 16, 9308-9319. 64. Lin, J.-D.; Wu, S.-T.; Li, Z.-H.; Du, S.-W., A Series of Novel Pb (II) or Pb (II)/M (II)(M= Ca and Sr) Hybrid Inorganic–Organic Frameworks Based on Polycarboxylic Acids with Diverse Pb– O–M (M= Pb, Ca and Sr) Inorganic Connectivities. CrystEngComm 2010, 12, 4252-4262. 65. Zhang, Y.; Wang, J.; Yan, X.; Liu, X.; Zhou, H.; Yuan, A., Syntheses, Structures, Thermal Stabilities and Gas Sorption Properties of Two Rod-Based Microporous Lead (II) Polycarboxylate Coordination Frameworks. Microporous Mesoporous Mater. 2014, 184, 15-20. 66. Ashbrook, S. E.; Dawson, D. M.; Seymour, V. R., Recent Developments in Solid-State NMR Spectroscopy of Crystalline Microporous Materials. Phys. Chem. Chem. Phys. 2014, 16, 8223-8242. 49

ACS Paragon Plus Environment

Chemistry of Materials

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

67. Martineau, C., NMR Crystallography: Applications to Inorganic Materials. Solid State Nucl. Magn. Reson. 2014, 63, 1-12. 68. Zhang, J.-P.; Liao, P.-Q.; Zhou, H.-L.; Lin, R.-B.; Chen, X.-M., Single-Crystal X-Ray Diffraction Studies on Structural Transformations of Porous Coordination Polymers. Chem. Soc. Rev. 2014, 43, 5789-5814. 69. Takamizawa, S.; Nakata, E.-i.; Saito, T.; Kojima, K., Structural Determination of Physisorbed Sites for CO2 and Ar Gases inside an Organometallic Framework. CrystEngComm 2003, 5, 411-413. 70. Takamizawa, S.; Takasaki, Y.; Miyake, R., Host–Guest Transformational Correlations for a Gas Inclusion Co-Crystal on Changing Gas Pressure and Temperature. Chem. Commun. 2009, 6625-6627. 71. Wriedt, M.; Sculley, J. P.; Yakovenko, A. A.; Ma, Y.; Halder, G. J.; Balbuena, P. B.; Zhou, H. C., Low-Energy Selective Capture of Carbon Dioxide by a Pre-Designed Elastic Single-Molecule Trap. Angew. Chem. Int. Ed. 2012, 124, 9942-9946. 72. Zhao, P.; Lampronti, G. I.; Lloyd, G. O.; Wharmby, M. T.; Facq, S. B.; Cheetham, A. K.; Redfern, S. A., Phase Transitions in Zeolitic Imidazolate Framework 7: The Importance of Framework Flexibility and Guest-Induced Instability. Chem. Mater. 2014, 26, 1767-1769. 73. Sotelo, J.; Woodall, C. H.; Allan, D. R.; Gregoryanz, E.; Howie, R. T.; Kamenev, K. V.; Probert, M. R.; Wright, P. A.; Moggach, S. A., Locating Gases in Porous Materials: Cryogenic Loading of Fuel-Related Gases into a Sc-Based Metal-Organic Framework under Extreme Pressures. Angew. Chem. Int. Ed. 2015, 127, 13530-13534. 74. He, P.; Lucier, B. E.; Terskikh, V. V.; Shi, Q.; Dong, J.; Chu, Y.; Zheng, A.; Sutrisno, A.; Huang, Y., Spies Within Metal-Organic Frameworks: Investigating Metal Centers Using Solid-State NMR. 50

ACS Paragon Plus Environment

Page 50 of 58

Page 51 of 58

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

Chemistry of Materials

J. Phys. Chem. C 2014, 118, 23728-23744. 75. Kong, X.; Deng, H.; Yan, F.; Kim, J.; Swisher, J. A.; Smit, B.; Yaghi, O. M.; Reimer, J. A., Mapping of Functional Groups in Metal-Organic Frameworks. Science 2013, 341, 882-885. 76. Rossini, A. J.; Zagdoun, A.; Lelli, M.; Canivet, J.; Aguado, S.; Ouari, O.; Tordo, P.; Rosay, M.; Maas, W. E.; Copéret, C., Dynamic Nuclear Polarization Enhanced Solid-State NMR Spectroscopy of Functionalized Metal-Organic Frameworks. Angew. Chem. 2012, 124, 127-131. 77. Gonzalez, J.; Devi, R. N.; Tunstall, D. P.; Cox, P. A.; Wright, P. A., Deuterium NMR Studies of Framework and Guest Mobility in the Metal–Organic Framework Compound MOF-5, Zn4O(O2CC6H4CO2)3. Microporous Mesoporous Mater. 2005, 84, 97-104. 78. Wang, W. D.; Lucier, B. E.; Terskikh, V. V.; Wang, W.; Huang, Y., Wobbling and Hopping: Studying Dynamics of CO2 Adsorbed in Metal–Organic Frameworks Via 17O Solid-State NMR. J. Phys. Chem. Lett. 2014, 5, 3360-3365. 79. Hoffmann, H. C.; Debowski, M.; Müller, P.; Paasch, S.; Senkovska, I.; Kaskel, S.; Brunner, E., Solid-State NMR Spectroscopy of Metal–Organic Framework Compounds (MOFs). Materials 2012, 5, 2537-2572. 80. Ibrahim, B.; Lucier, B. E.; Xu, J.; He, P.; Huang, Y., Investigating Adsorption of Organic Compounds in Metal-Organic Framework MIL-53. Can. J. Chem. 2015, 93, 960-969. 81. Gul-E-Noor, F.; Jee, B.; Pöppl, A.; Hartmann, M.; Himsl, D.; Bertmer, M., Effects of Varying Water Adsorption on a Cu3 (BTC)2 Metal–Organic Framework (MOF) as Studied by 1H and

13

C

Solid-State NMR Spectroscopy. Phys. Chem. Chem. Phys. 2011, 13, 7783-7788. 82. Sutrisno, A.; Huang, Y., Solid-State NMR: A Powerful Tool for Characterization of Metal– Organic Frameworks. Solid State Nucl. Magn. Reson. 2013, 49, 1-11. 83. Xu, J.; Terskikh, V. V.; Chu, Y.; Zheng, A.; Huang, Y., Mapping out Chemically Similar, 51

ACS Paragon Plus Environment

Chemistry of Materials

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

Crystallographically Nonequivalent Hydrogen Sites in Metal–Organic Frameworks by

Page 52 of 58

1

H

Solid-State NMR Spectroscopy. Chem. Mater. 2015, 27, 3306-3316. 84. Volkringer, C.; Popov, D.; Loiseau, T.; Férey, G.; Burghammer, M.; Riekel, C.; Haouas, M.; Taulelle, F., Synthesis, Single-Crystal X-Ray Microdiffraction, and NMR Characterizations of the Giant Pore Metal-Organic Framework Aluminum Trimesate MIL-100. Chem. Mater. 2009, 21, 5695-5697. 85. Xu, J.; Lucier, B. E.; Sinelnikov, R.; Terskikh, V. V.; Staroverov, V. N.; Huang, Y., Monitoring and Understanding the Paraelectric-Ferroelectric Phase Transition in the Metal-Organic Framework [NH4][M(HCOO)3] by Solid-State NMR Spectroscopy. Chem. Eur. J. 2015, 21, 14348-14361. 86. Kong, X.; Scott, E.; Ding, W.; Mason, J. A.; Long, J. R.; Reimer, J. A., CO2 Dynamics in a Metal–Organic Framework with Open Metal Sites. J. Am. Chem. Soc. 2012, 134, 14341-14344. 87. Lin, L. C.; Kim, J.; Kong, X.; Scott, E.; McDonald, T. M.; Long, J. R.; Reimer, J. A.; Smit, B., Understanding CO2 Dynamics in Metal–Organic Frameworks with Open Metal Sites. Angew. Chem. Int. Ed. 2013, 125, 4506-4509. 88. Lucier, B. E.; Chan, H.; Zhang, Y.; Huang, Y., Multiple Modes of Motion: Realizing the Dynamics of CO Adsorbed in M-MOF-74 (M= Mg, Zn) by Using Solid‐State NMR Spectroscopy. Eur. J. Inorg. Chem. 2016, 2016, 2017-2024. 89. Zhang, Y.; Lucier, B. E.; Huang, Y., Deducing CO2 Motion, Adsorption Locations and Binding Strengths in a Flexible Metal–Organic Framework without Open Metal Sites. Phys. Chem. Chem. Phys. 2016, 18, 8327-8341. 90. Fulmer, G. R.; Miller, A. J.; Sherden, N. H.; Gottlieb, H. E.; Nudelman, A.; Stoltz, B. M.; Bercaw, J. E.; Goldberg, K. I., NMR Chemical Shifts of Trace Impurities: Common Laboratory Solvents, Organics, and Gases in Deuterated Solvents Relevant to the Organometallic Chemist. 52

ACS Paragon Plus Environment

Page 53 of 58

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

Chemistry of Materials

Organometallics 2010, 29, 2176-2179. 91. Cory, D.; Ritchey, W., Suppression of Signals from the Probe in Bloch Decay Spectra. J. Magn. Reson. 1988, 80, 128-132. 92. O'Dell, L. A., The WURST Kind of Pulses in Solid-State NMR. Solid State Nucl. Magn. Reson. 2013, 55, 28-41. 93. O’Dell, L. A.; Schurko, R. W., QCPMG Using Adiabatic Pulses for Faster Acquisition of Ultra-Wideline NMR Spectra. Chem. Phys. Lett. 2008, 464, 97-102. 94. Tang, J. A.; O’Dell, L. A.; Aguiar, P. M.; Lucier, B. E.; Sakellariou, D.; Schurko, R. W., Application of Static Microcoils and WURST Pulses for Solid-State Ultra-Wideline NMR Spectroscopy of Quadrupolar Nuclei. Chem. Phys. Lett. 2008, 466, 227-234. 95. O’Dell, L. A.; Rossini, A. J.; Schurko, R. W., Acquisition of Ultra-Wideline NMR Spectra from Quadrupolar Nuclei by Frequency Stepped WURST–QCPMG. Chem. Phys. Lett. 2009, 468, 330-335. 96. Bhattacharyya, R.; Frydman, L., Quadrupolar Nuclear Magnetic Resonance Spectroscopy in Solids Using Frequency-Swept Echoing Pulses. J. Chem. Phys. 2007, 127, 194503(1)-194503(8). 97. Kupce, Ē.; Freeman, R., Stretched Adiabatic Pulses for Broadband Spin Inversion. J. Magn. Reson., Series A 1995, 117, 246-256. 98. Dybowski, C.; Neue, G., Solid State

207

Pb NMR Spectroscopy. Prog. Nucl. Magn. Reson.

Spectrosc. 2002, 41, 153-170. 99. Mennitt, P. G.; Shatlock, M. P.; Bartuska, V. J.; Maciel, G. E., Cadmium-113 NMR Studies of Solid Cadmium (II) Complexes. J. Phys. Chem. 1981, 85, 2087-2091. 100.

Lee, G. S.; Fisher, K. J.; Vassallo, A. M.; Hanna, J. V.; Dance, I. G., Solid-State

Cadmium-113

NMR

of

Three

Structural

Isomers

of

Hexadecakis

53

ACS Paragon Plus Environment

(Benzenethiolato)

Chemistry of Materials

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

Tetrasulfidodecacadmate (4-). Inorg. Chem. 1993, 32, 66-72. 101.

Eichele, K., WSolids1, Version 1.20. 15. Universität Tübingen: Tübingen, Germany 2011.

102.

Vold, R. L.; Hoatson, G. L., Effects of Jump Dynamics on Solid State Nuclear Magnetic

Resonance Line Shapes and Spin Relaxation Times. J. Magn. Reson. 2009, 198, 57-72. 103.

Beeler, A. J.; Orendt, A. M.; Grant, D. M.; Cutts, P. W.; Michl, J.; Zilm, K. W.; Downing, J.

W.; Facelli, J. C.; Schindler, M. S.; Kutzelnigg, W., Low-Temperature Carbon-13 Magnetic Resonance in Solids. 3. Linear and Pseudolinear Molecules. J. Am. Chem. Soc. 1984, 106, 7672-7676. 104.

Bruker-AXS, SAINT version 2013.8, 2013, Bruker-AXS, Madison, WI 53711, USA.

105.

Bruker-AXS, SADABS version 2012.1, 2012, Bruker-AXS, Madison, WI 53711, USA.

106.

Sheldrick, G. M., Crystal Structure Refinement with Shelxl. Acta Cryst. 2015, 71, 3-8.

107.

Sheldrick, G., A Short History of Shelx. Acta Cryst. 2008, 64, 112-122.

108.

Berglund, M.; Wieser, M. E., Isotopic Compositions of the Elements 2009 (IUPAC

Technical Report). Pure Appl. Chem. 2011, 83, 397-410. 109.

Zhang, Y.; Lucier, B. E.; Huang, Y., Deducing CO2 Motion, Adsorption Locations and

Binding Strengths in a Flexible Metal–Organic Framework without Open Metal Sites. Phys. Chem. Chem. Phys. 2016, 18, 8327-8341. 110. Torrisi, A.; Mellot-Draznieks, C.; Bell, R. G., Impact of Ligands on CO2 Adsorption in Metal-Organic Frameworks: First Principles Study of the Interaction of CO2 with Functionalized Benzenes. I. Inductive Effects on the Aromatic Ring. J. Chem. Phys. 2009, 130, 194703-194703. 111. Sutrisno, A.; Terskikh, V. V.; Shi, Q.; Song, Z.; Dong, J.; Ding, S. Y.; Wang, W.; Provost, B. R.; Daff, T. D.; Woo, T. K., Characterization of Zn-Containing Metal-Organic Frameworks by Solid-State

67

Zn NMR Spectroscopy and Computational Modeling. Chem. Eur. J. 2012, 18, 54

ACS Paragon Plus Environment

Page 54 of 58

Page 55 of 58

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

Chemistry of Materials

12251-12259. 112. Xu, J.; Terskikh, V. V.; Huang, Y., Resolving Multiple Non-Equivalent Metal Sites in Magnesium-Containing Metal-Organic Frameworks by Natural Abundance 25Mg Solid-State NMR Spectroscopy. Chem. –Eur. J. 2013, 19, 4432-4436. 113. Haouas, M.; Volkringer, C.; Loiseau, T.; Férey, G.; Taulelle, F., The Extra-Framework Sub-Lattice of the Metal-Organic Framework MIL-110: A Solid-State NMR Investigation. Chem. Eur. J. 2009, 15, 3139-3146. 114. Haouas, M.; Volkringer, C.; Loiseau, T.; Férey, G.; Taulelle, F., In Situ NMR, Ex Situ XRD and SEM Study of the Hydrothermal Crystallization of Nanoporous Aluminum Trimesates MIL-96, MIL-100, and MIL-110. Chem. Mater. 2012, 24, 2462-2471. 115. Reimer, N.; Reinsch, H.; Inge, A. K.; Stock, N., New Al-MOFs Based on Sulfonyldibenzoate Ions: A Rare Example of Intralayer Porosity. Inorg. Chem. 2014, 54, 492-501. 116. Venkataraman, D.; Gardner, G. B.; Lee, S.; Moore, J. S., Zeolite-Like Behavior of a Coordination Network. J. Am. Chem. Soc. 1995, 117, 11600-11601. 117. Schaate, A.; Roy, P.; Godt, A.; Lippke, J.; Waltz, F.; Wiebcke, M.; Behrens, P., Modulated Synthesis of Zr-Based Metal-Organic Frameworks: From Nano to Single Crystals. Chem. Eur. J. 2011, 17, 6643-6651. 118. Modrow, A.; Zargarani, D.; Herges, R.; Stock, N., The First Porous MOF with Photoswitchable Linker Molecules. Dalton Trans. 2011, 40, 4217-4222. 119. Hisaki, I.; Senga, H.; Shigemitsu, H.; Tohnai, N.; Miyata, M., Construction of 1D π-Stacked Superstructures with Inclusion Channels through Symmetry-Decreasing Crystallization of Discotic Molecules of C3 Symmetry. Chem. Eur. J. 2011, 17, 14348-14353. 120.

Huang, W.; Jiang, J.; Wu, D.; Xu, J.; Xue, B.; Kirillov, A. M., A Highly Stable Nanotubular 55

ACS Paragon Plus Environment

Chemistry of Materials

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

MOF Rotator for Selective Adsorption of Benzene and Separation of Xylene Isomers. Inorg. Chem. 2015, 54, 10524-10526. 121.

Banerjee, D.; Simon, C. M.; Plonka, A. M.; Motkuri, R. K.; Liu, J.; Chen, X.; Smit, B.;

Parise, J. B.; Haranczyk, M.; Thallapally, P. K., Metal–Organic Framework with Optimally Selective Xenon Adsorption and Separation. Nat. Commun. 2016, 7, ncomms11831(1)ncomms11831(7). 122.

Wu, H.; Simmons, J. M.; Srinivas, G.; Zhou, W.; Yildirim, T., Adsorption Sites and

Binding Nature of CO2 in Prototypical Metal-Organic Frameworks: A Combined Neutron Diffraction and First-Principles Study. J. Phys. Chem. Lett. 2010, 1, 1946-1951. 123.

Yao, Y.; Nijem, N.; Li, J.; Chabal, Y. J.; Langreth, D. C.; Thonhauser, T., Analyzing the

Frequency Shift of Physiadsorbed CO2 in Metal Organic Framework Materials. Phys. Rev. B 2012, 85, 064302(1)-064302(8). 124.

Poloni, R.; Smit, B.; Neaton, J. B., CO2 Capture by Metal–Organic Frameworks with Van

Der Waals Density Functionals. J. Phys. Chem. A 2012, 116, 4957-4964. 125.

Vaidhyanathan, R.; Iremonger, S. S.; Shimizu, G. K.; Boyd, P. G.; Alavi, S.; Woo, T. K.,

Direct Observation and Quantification of CO2 Binding within an Amine-Functionalized Nanoporous Solid. Science 2010, 330, 650-653.

56

ACS Paragon Plus Environment

Page 56 of 58

Page 57 of 58

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

Chemistry of Materials

57

ACS Paragon Plus Environment

Chemistry of Materials

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

TOC image

58

ACS Paragon Plus Environment

Page 58 of 58