Metal–Organic Frameworks with Reduced Hydrophilicity for

Aug 2, 2018 - Postcombustion CO2 capture from wet flue gas is a daunting challenge that metal–organic frameworks (MOFs) based adsorbents need to ...
0 downloads 0 Views 956KB Size
Subscriber access provided by University of Sussex Library

Article

Metal–Organic Frameworks with Reduced Hydrophilicity for Postcombustion CO2 Capture from Wet Flue Gas Yuxiang Wang, Zhigang Hu, Tanay Kundu, Youdong Cheng, Jinqiao Dong, Yuhong Qian, Linzhi Zhai, and Dan Zhao ACS Sustainable Chem. Eng., Just Accepted Manuscript • DOI: 10.1021/ acssuschemeng.8b02173 • Publication Date (Web): 02 Aug 2018 Downloaded from http://pubs.acs.org on August 6, 2018

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

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

Page 1 of 26 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

ACS Sustainable Chemistry & Engineering

Metal–Organic Frameworks with Reduced Hydrophilicity for Postcombustion CO2 Capture from Wet Flue Gas Yuxiang Wang, Zhigang Hu, Tanay Kundu, Youdong Cheng, Jinqiao Dong, Yuhong Qian, Linzhi Zhai, and Dan Zhao* Department of Chemical and Biomolecular Engineering, National University of Singapore, 4 Engineering Drive 4, 117585, Singapore *Email Address: [email protected]

ABSTRACT: Postcombustion CO2 capture from wet flue gas is a daunting challenge that metal–organic frameworks (MOFs) based adsorbents need to address, because the moisture in the gas stream may not only hydrolyze the coordination bonds of MOFs but also be competitively adsorbed in MOFs leading to compromised CO2 capture performance. In this study, two isostructural water-stable MOFs decorated with alkyl groups, namely UiO-66(Zr)(OAc)2 and UiO-66(Zr)-(OPr)2, are synthesized from UiO-66(Zr)-(OH)2 via a facile postsynthetic esterification strategy and evaluated for their water affinity and CO2 capture performance. The increased water contact angle and reduced water vapor capacity at 60% relative humidity indicate the positive role of propionyl group of UiO-66(Zr)-(OPr)2 in reducing material hydrophilicity. When the adsorbent beds are not fully saturated with water,

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

breakthrough experiments using simulated wet flue gas reveal that UiO-66(Zr)-(OPr)2 possesses a CO2/N2 adsorption selectivity of 76.6, which is 229% of that of UiO-66(Zr)-(OH)2 (33.4). Our study successfully demonstrates a scalable material modification approach to engineer MOF adsorbents toward practical CO2 capture processes under wet conditions.

Keywords: Postcombustion CO2 capture, Pressure swing adsorption, Metal-organic frameworks, Post-synthetic modification, Wet gas breakthrough

INTRODUCTION Global warming because of escalating anthropogenic greenhouse gas emission has aroused worldwide concerns. Owing to the slow-pacing development of sustainable energy supplies, CO2 capture and utilization is considered as a promising solution for the dilemma between ascending energy demand and pressing deterioration of environment.1-2 Despite being the most commonly applied method for CO2 capture in industry, amine scrubbing suffers from equipment corrosion, solvent loss, amine degradation, and high energy consumption to regenerate strongly absorbed CO2 from the absorbents.3-5 On the contrary, adsorption-based CO2 capture techniques such as pressure/temperature swing adsorption are attracting increasing interests due to their lower operational cost, less toxicity, and milder corrosivity.1,6 Ideally, adsorbents used for CO2 capture should possess high working capacity (defined as the CO2 uptake capacity difference between adsorption and desorption stages in the process), high gas selectivity of CO2 over N2, decent chemical resistivity and mechanical strength, minimalized water affinity, and low material costs.7-9 Among many adsorbents being studied, metal–organic frameworks (MOFs), a category of porous crystalline materials constructed from inorganic clusters and organic bridging ligands,

ACS Paragon Plus Environment

Page 2 of 26

Page 3 of 26 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

ACS Sustainable Chemistry & Engineering

have demonstrated enormous potential in gas separation processes, especially in CO2 capture.2,1013

Because of the hygroscopic nature of CO2 binding sites in MOFs (e.g., unsaturated open metal

sites, polar functional groups, and metal clusters), an inevitable challenge that many MOF adsorbents are facing is the competitive adsorption of water over CO2 in wet flue gas which typically contains 5−10 vol% of water vapor. This competitive adsorption of water will not only compromise the adsorbent working capacity but also increase the energy required for adsorbent regeneration.14-16 In addition, the hydrothermal stability of MOFs is another issue handicapping the practical applications of MOF-based CO2 capture.17 For example, the CO2 uptake capacities of MOF-74(Mg) and SIFSIX-3-Ni are 5.34 and 2.48 mmol g−1, respectively, at 303 K and 0.15 atm under dry conditions. Nevertheless, these values shrink significantly to 1.55 and 1.73 mmol g-1, respectively, when the 15/85 CO2/N2 gas mixture at 1 atm is saturated with moisture at room temperature.15 This striking decline in CO2 adsorption capacity is likely due to the competitive adsorption of water along with the structural evolutions of the MOFs, namely, partial structure collapse in MOF-74(Mg) and gradual phase transition in SIFSIX-3-Ni.17-18 To overcome the adverse effects of moisture on CO2 adsorption in MOFs, enhancing CO2/H2O adsorption selectivity by the introduction of chemisorption sites such as amine groups into the nanospace of MOFs has been extensively studied.19-23 A series of diamines have been loaded into MOFs with open metal sites such as M2(dobpdc) (M = Mg, Fe, Co, Zn, Mn, dobpdc4- = 4,4’dioxidobiphenyl-3,3’-dicarboxylate)

and

Mg2(dobdc)

(dobdc4-

=

2,5-dioxido-1,4-

benzenedicarboxylate) via the coordination between amino groups and metal cations, resulting in MOFs capable of performing effective CO2 capture at elevated temperatures in the presence of moisture.19,24-26 The second general strategy is endowing hydrophobicity to MOFs to weaken the interactions between MOFs and water so that the competition adsorption of water can be

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

relieved.27 For instance, alkyl chains and perfluoroalkyl groups have be immobilized into MOFs by postsynthetic acylation, thiol-ene click reaction, or liquid phase chemisorption to provide excellent hydrophobicity.28-30 In addition, hybridizing MOFs with hydrophobic polymers can also modify the surface properties of the MOFs.31 Apart from the commonly used polydimethysiloxane (PDMS) as the protective hydrophobic layer, reduced graphene oxide (rGO) and microporous organic networks (MONs) have also been employed to prepare MOF/polymer composites with enhanced water stability.32-35 Recently, multivariate (MTV) MOFs incorporating hydrophobic ligands have been demonstrated to counteract the moisture interference effectively in MOF-based CO2 capture from wet flue gas. MTV zeolitic imidazole frameworks (ZIFs) containing 2-mthylimidazolate and a series of benzimidazolate derivatives exhibit equally effective CO2 separation from N2 in both the presence and absence of water.36 In a recent study, our group synthesized MTV UiO-66(Zr) incorporating aminoterephthalate (NH2BDC) and tetrafluoroterephthalate (TFBDC) by a triphasic modulated hydrothermal (MHT) synthesis method.37 Dynamic breakthrough experiments show that the MTV MOF UiO-66(Zr)NH2-F4-0.53 only loses 30% of its CO2 uptake capacity in simulated wet flue gas mixture at 298 K, outperforming the almost 100% loss of zeolite materials and other hydrophilic adsorbents under similar conditions. This study highlights the importance of engineering water affinity of MOFs for moisture-resistant CO2 capture applications. Opt-UiO-66(Zr)-(OH)2 is a water-stable Zr MOF developed in our lab with a high CO2 uptake capacity (2.52 mmol g-1 in a 15/85 CO2/N2 gas mixture at an equilibrium pressure of 1.3 bar, 298 K), outstanding CO2/N2 adsorption selectivity (~45 based on dynamic column breakthrough experiments), and excellent hydrothermal stability.38 However, up to 85% loss in CO2 uptake capacity was observed when this MOF was saturated with moisture under the atmosphere of 80%

ACS Paragon Plus Environment

Page 4 of 26

Page 5 of 26 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

ACS Sustainable Chemistry & Engineering

relative humidity (RH), which undesirably impedes its practical CO2 capture applications. To overcome this drawback, we hypothesize that the hygroscopic nature of this MOF could be altered by converting its hydrophilic hydroxyl groups to less polar ester groups along with the introduction of aliphatic chains so that the adverse effects of moisture on CO2 adsorption might be alleviated. Herein, we show that postsynthetic esterification of UiO-66(Zr)-(OH)2 using aliphatic carboxylic anhydride can indeed reduce the hydrophilicity of the MOF precursor without damaging its intrinsic stability. As demonstrated by the breakthrough experiments using adsorption beds which are not saturated with water beforehand, one of the esterified MOFs, UiO66(Zr)-(OPr)2, can capture CO2 from simulated wet flue gas with a higher CO2/N2 co-adsorption selectivity than that of UiO-66(Zr)-(OH)2. Our proof-of-concept study suggests the promising direction of water affinity engineering in MOFs for enhanced CO2 capture performance under wet conditions.

EXPERIMENTAL SECTION Modulated Hydrothermal (MHT) Synthesis of UiO-66(Zr)-(OH)2. UiO-66(Zr)-(OH)2 was synthesized according to the previously reported method.38 Briefly, 2,5-dihydroxyterephthalic acid (H4DOBDC, 2 g, ~10 mmol) and ZrOCl2·8H2O (3.4 g, ~10.4 mmol) were added in a 100 mL water/acetic acid (20/30, v/v) mixture and heated under reflux (~105 °C) for 1 day to afford the dark khaki powder. The powder was washed with water for three times before soaking in water and methanol each for 3 days at room temperature, during which the solvent was refreshed every day. After the final removal of methanol by decanting, the sample was activated under vacuum at 120 °C overnight for further modifications and characterizations.

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

Postsynthetic Esterification of UiO-66(Zr)-(OH)2. A suspension of activated UiO-66(Zr)(OH)2 (100 mg, ~0.65 mmol) in 5 mL of aliphatic carboxylic anhydride [acetic anhydride for UiO-66(Zr)-(OAc)2 and propionic anhydride for UiO-66(Zr)-(OPr)2] was heated to 55 °C before adding three drops of concentrated H2SO4. After 5 hours of heating, dark khaki powder was isolated by centrifugation and washed with tetrahydrofuran (THF) for three times. The powder was further soaked in THF for 3 days, during which the solvent was refreshed every day. Finally, the sample was activated under vacuum at 120 °C overnight for further characterizations. Breakthrough Experiments. The breakthrough experiments were conducted using a homebuilt set up shown in Scheme S1. Gas composition at the exit of the column was determined by a mass spectrometer (Hiden QGA quantitative gas analysis system). The flow rate of each component gas was calculated by an internal Ar flow reference that has a fixed flow rate of 5 ± 0.01 sccm (standard cubic centimeter per minute). Before the breakthrough experiments, CO2/N2 (15 ± 1) / (85 ± 1) binary mixture gas with a flow rate of 5 ± 0.1 sccm was introduced through the bypass line with various resistances (pressure drop) controlled by a needle valve. The corresponding gas residence time was calculated for the calibration of system dead volume. A 7 cm stainless steel column (internal diameter: 0.46 cm) filled with MOF powders (~0.6 g depending on the adsorbent density) was activated by purging a constant He flow (8 ± 0.1 sccm) through the column at 60 °C for 1 day until no solvent or moisture signal could be detected by the mass spectrometer. For dry gas breakthrough experiments, a (15 ± 1)/(85 ± 1) mixture of CO2 and N2 with a total flow rate of 5 ± 0.1 sccm was stabilized for 15 minutes before being introduced into the column. As for experiments of wet gas breakthrough in the MOF-packed columns, a (15 ± 1)/(85 ± 1) CO2/N2 mixture with a relative humidity (RH) around 60% was directly introduced into the system after stabilization for 15 minutes. Pre-saturation of packed

ACS Paragon Plus Environment

Page 6 of 26

Page 7 of 26 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

ACS Sustainable Chemistry & Engineering

columns with water vapor was conducted by purging a (15 ± 1)/(85 ± 1) He/N2 mixture (RH ~60%) through the column after activation described above. Mild activation was conducted by purging the packed column with a constant He flow (8 ± 0.1 sccm) at room temperature for 3 hours. Methods to calculate gas co-adsorption capacity and selectivity based on breakthrough experiments are detailed in the supporting information.

RESULTS AND DISCUSSION Structure and Porosity Characterization. There are two paths to introduce alkyl chains into the framework of UiO-66-(OH)2 by esterification of the hydroxyl groups (Scheme 1). The first path can be regarded as a direct approach, namely, synthesizing esterified ligand first and subsequently using the modified ligand to grow corresponding MOFs by solvothermal synthesis.39 The second path is the postsynthetic approach where esterification to immobilize alkyl chains is conducted after MOF formation. The direct approach was initially adopted because this method was expected to afford a higher esterification ratio than the MOFs obtained by postsynthetic approach which may have undesirable diffusion limitations of the reactants. Nevertheless, as shown in Figure S1, the MOF prepared by this approach displays a Fouriertransform infrared spectroscopy (FTIR) spectrum similar to that of UiO-66(Zr)-(OH)2 with no band at about 1750 cm-1 from ester carbonyl groups, suggesting that the ester bonds of the ligands were hydrolyzed during the process of solvothermal synthesis. Therefore, the postsynthetic approach was used instead in this study.

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

Scheme 1. Schematic illustration of the strategy combing modulated hydrothermal (MHT) synthesis and postsynthetic modification to prepare esterified MOFs. The MOF precursor UiO-66(Zr)-(OH)2 was synthesized based on a previously reported protocol of modulated hydrothermal (MHT) synthesis.38,40-41 The phase purity and quality of the MOF were examined by powder X-ray diffraction (PXRD, Figure 1a), thermogravimetric analysis (TGA, Figure S2), and Ar sorption at 87 K (Figure 2a), all confirming the successful repetition of the reported method. Calculation based on the TGA data shows that about 1.44 linkers are missing for each Zr6 cluster, leaving vacant sites capped by acetate anions as indicated by the proton signal of acetyl at around 1.7 ppm in the 1H-NMR spectra (Figure S3).42 After activation, UiO-66(Zr)-(OH)2 was subsequently added into aliphatic carboxylic anhydride and the mixture was heated to about 55 °C for 10 hours under the catalysis of concentrated sulfuric acid. The PXRD patterns of esterified MOFs are consistent with that of the precursor MOF (Figure 1a). In addition, there is no obvious change in the particle size or morphology of the MOFs before and after esterification (Figure S4), confirming the structural integrity of UiO66-Zr-(OH)2 in the highly corrosive reaction environment of postsynthetic esterification. For the MOFs treated with acetic anhydride and propionic anhydride, which are denoted as UiO-66(OAc)2 and UiO-66-(OPr)2, respectively, new FTIR peaks at ~1740 cm-1 corresponding to the

ACS Paragon Plus Environment

Page 8 of 26

Page 9 of 26 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

ACS Sustainable Chemistry & Engineering

ester carbonyl groups can be observed (Figure 1b).43 Besides, the intensity of C−O stretching in the ligands at 1231 cm-1 decreases noticeably compared to the intensity of carbonyl groups in the carboxylate, suggesting that part of the hydroxyl groups have been converted during the esterification process.43 The FTIR spectrum of UiO-66(Zr)-(OiBu)2 obtained by treating the MOF precursor with isobutyric anhydride, however, does not exhibit the characteristic band of ester carbonyl groups. Instead, there is a new band at about 1646 cm-1 indicating the presence of uncoordinated aromatic acid, which can be attributed to the partial decomposition of the MOF.44 This unsuccessful esterification using isobutyric anhydride is likely due to the steric effect of the reactants and diffusion limitations in the confined nanospace of UiO-66(Zr)-(OH)2. 1H-NMR spectra of the digested esterified MOFs were collected to estimate the conversion ratio of phenolic hydroxyl groups (Figure S3). For digested UiO-66(Zr)-(OAc)2, the acetyl proton signal has an integration area of ca. 3.7 with reference to the phenyl proton signal (integration area assumed to be 1), which suggests 100% esterification yield of the phenolic hydroxyl groups if the capping acetates in the MOF are taken into consideration. On the other hand, the acetyl methyl proton signal almost disappears for digested UiO-66(Zr)-(OPr)2, suggesting that the capping acetates were replaced by propionates after the postsynthetic modification. Assuming that the amount of the defects remained unchanged, the esterification yield is estimated to be 44% for UiO-66(Zr)-(OPr)2.

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

(a)

UiO-66(Zr)-(OiBu)2

(b)

UiO-66(Zr)-(OPr)2

Transmittance / a.u.

UiO-66(Zr)-(OAc)2 UiO-66(Zr)-(OH)2

Intensity / a.u.

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

Page 10 of 26

1740 cm-1 1646 cm-1 UiO-66(Zr)-(OH)2 UiO-66(Zr)-(OAc)2 UiO-66(Zr)-(OPr)2

1231 cm-1

UiO-66(Zr)-(OiBu)2

5

10

15

20

25

30

35

40

2000

1800

1600

1400

Wavenumber / cm

2 Theta / degree

1200

1000

-1

Figure 1. (a) PXRD patterns and (b) FTIR spectra of UiO-66(Zr)-(OH)2 and its esterified derivatives. Ar sorption tests at 87 K were conducted to probe the changes in the internal texture such as porosity and pore size distribution of the MOFs after postsynthetic modification. As shown in Figure 2a, the MOFs exhibit typical type I isotherms with minor desorption hystereses at high P/P0 values. The BET surface area of UiO-66(Zr)-(OH)2 drops from 929.7 m2 g-1 to 552.0 m2 g-1 for UiO-66(Zr)-(OAc)2 and 574 m2 g-1 for UiO-66(Zr)-(OPr)2. The modified MOFs possess comparable surface area even though propionyl groups in UiO-66(Zr)-(OPr)2 are bulkier than acetyl groups in UiO-66(Zr)-(OAc)2, which is possibly due to the lower alkyl chain loading in UiO-66(Zr)-(OPr)2 indicated by 1H-NMR. According to the pore size distribution (Figure 2b) calculated by non-local density functional theory (NLDFT) model, the esterification process reduces the total amount of pores, especially those with pore sizes around 5.3 Å, suggesting that the alkyl carbonyl moieties mainly reside in the octahedral pores of the MOFs.

ACS Paragon Plus Environment

(a)

(b) 250 200

Vads / cm3stp g-1

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

ACS Sustainable Chemistry & Engineering

150 100 UiO-66(Zr)-(OH)2 UiO-66(Zr)-(OAc)2

50

UiO-66(Zr)-(OPr)2

0 0.0

0.2

0.4

0.6

0.8

1.0

Incremental Pore Volume / cm3 g-1 Å-1

Page 11 of 26

0.6

UiO-66(Zr)-(OH)2 UiO-66(Zr)-(OAc)2 UiO-66(Zr)-(OPr)2

0.4

0.2

0.0 4

Relative Pressure P/P0

6

8

Pore Width / Å

Figure 2. (a) Ar sorption isotherms at 87 K and (b) pore size distributions of UiO-66(Zr)-(OH)2, UiO-66(Zr)-(OAc)2, and UiO-66(Zr)-(OPr)2. Assessment of Water Affinity of MOFs. There are two potential water adsorption sites in UiO-66(Zr)-(OH)2: zirconium-oxo clusters and hydroxyl groups from the ligands. Postsynthetic esterification is expected to convert the hydroxyl adsorption sites to less hydrophilic moieties to alleviate the competitive sorption of water during CO2 capture processes. Besides, replacing the node-capping acetates by propionate may also weaken the water affinity of UiO-66(Zr)-(OPr)2. Water vapor sorption at 293 K and water contact angle measurement were performed to assess water affinity in the esterified MOFs and the MOF precursor. As shown in Figure 3a, the initial slopes of water isotherms of the three MOFs are close to each other, indicating small changes in the thermodynamic water affinity of the MOFs after the introduction of alkyl chains. These results suggest that the introduced alkyl chains are not long enough to block the exposed hydrophilic metal clusters, which are probably the major water adsorption sites at low pressure ranges. At higher RHs (e.g., 60%), UiO-66(Zr)-(OH)2 exhibits a higher water adsorption capacity (15.5 mmol g-1) than that of UiO-66(Zr)-(OAc)2 (10.7 mmol g-1) and UiO-66(Zr)-(OPr)2 (9.0 mmol g-1), which is likely due to the larger porosity and the presence of hydroxyl sorption sites in the precursor MOF. Interestingly, although UiO-66(Zr)-(OPr)2 has a surface area

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

comparable to that of UiO-66(Zr)-(OAc)2, the water sorption capacity of the former is noticeably lower than that of the latter, suggesting that the less polar propionyl group can endow the MOF with weaker hydrophilicity. Furthermore, UiO-66(Zr)-(OPr)2 exhibits larger contact angles (mean value, 73.6°) than that of UiO-66(Zr)-(OH)2 (mean value, 36.4°) and UiO-66(Zr)-(OAc)2 (mean value, 36.3°) as shown in Figure 3b, 3c and Figure S5, confirming the effective role of propionyl groups in reducing the hydrophilicity of the MOF.

(a) 15 Gas Uptake / mmol g-1

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

(b)

12

9

6 UiO-66(Zr)-(OH)2

3

(c)

UiO-66(Zr)-(OAc)2 UiO-66(Zr)-(OPr)2

0 0.0

0.2

0.4

0.6

0.8

Relative Pressure P/P0

Figure 3. (a) Water vapor sorption isotherms at 293 K of UiO-66(Zr)-(OH)2, UiO-66(Zr)(OAc)2, and UiO-66(Zr)-(OPr)2 (solid symbols, adsorption; open symbols, desorption). Snapshots of water contact angle measurements of (b) UiO-66(Zr)-(OPr)2 and (c) UiO-66(Zr)(OH)2. CO2 Capture Performance Based on Isotherms. Low pressure CO2 and N2 sorption properties of the esterified MOFs were studied by measuring gas sorption isotherms at 298 K and 273 K (Figure 4a, Figure S6). At 298 K, UiO-66(Zr)-(OH)2 shows a CO2 uptake capacity of 1.92 mmol g−1 at 0.15 bar (0.15 bar is the partial pressure of CO2 in flue gas). This value is reasonably lower than the reported one (2.50 mmol g-1) because the BET surface area of UiO-66(Zr)-(OH)2 (929.7 m2 g-1) has not been optimized to the value in the literature (1230 m2 g-1).38 Although the

ACS Paragon Plus Environment

Page 12 of 26

Page 13 of 26 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

ACS Sustainable Chemistry & Engineering

CO2 uptake capacities of UiO-66(Zr)-(OAc)2 and UiO-66(Zr)-(OPr)2 are 1.28 mmol g-1 and 1.19 mmol g-1, 33% and 38% lower than that of their precursor, respectively, these values are still among the best of all UiO-66 type materials.45-46 The decline in CO2 uptake capacity of the MOFs can be attributed to lower surface area and fewer polar hydroxyl groups in the modified MOFs. To probe the interactions between CO2 molecules and MOFs, the CO2 isosteric heat of adsorption (Qst) of the MOFs were calculated using the Clausius-Clapeyron equation based on isotherms at 298 K and 273 K (Figure 4b and Figure S7).47 As expected, UiO-66(Zr)-(OH)2 exhibits a zero-coverage CO2 Qst of 34.5 kJ mol-1, which is comparable to the previously reported value (33 kJ mol-1).38 Although the zero-coverage CO2 Qst of UiO-66(Zr)-(OPr)2 is close to that of its precursor, UiO-66(Zr)-(OAc)2 possesses a larger zero-coverage CO2 Qst of about 36.2 kJ mol-1. We presume that this slightly intensified CO2 interaction with UiO-66(Zr)(OAc)2 may result from the enhanced van der Waals forces experienced by CO2 molecules due to the shrinkage of pore size in the MOF, while the loading of propionyl groups in UiO-66(Zr)(OPr)2 is not high enough to manifest this effect.48-50 Ideal adsorbed solution theory (IAST) was employed to predict the ideal CO2/N2 separation performance of the MOFs in this study for a 15/85 CO2/N2 gas mixture at 298 K. As shown in Figure 4c, UiO-66(Zr)-(OH)2 has an IAST CO2/N2 adsorption selectivity of 66 at 298 K and 1 bar, which falls between the previously reported values (34.2 and 105).38,40 As for the esterified MOFs, UiO-66(Zr)-(OAc)2 and UiO66(Zr)-(OPr)2 exhibit ideal CO2/N2 selectivities of 42.5 and 39.1, respectively, which are lower than that of UiO-66(Zr)-(OH)2 but still higher than most Zr MOFs such as UiO-66(Zr) (19.4),51 UiO-66(Zr)-NH2 (32.3),52 and UiO-66(Zr)-(COOH)2 (35.4).40

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

(a)

5

Gas Uptake / mmol g-1

4 3 2 UiO-66(Zr)-(OH)2

1

UiO-66(Zr)-(OAc)2 UiO-66(Zr)-(OPr)2

0 0

20

40

60

80

100

Absolute Pressure / kPa

(b)

40 35

Qst / kJ mol-1

30 25 20 UiO-66(Zr)-(OH)2 UiO-66(Zr)-(OAc)2

15

UiO-66(Zr)-(OPr)2

10 0

1

2

3

4

5

-1

Gas Adsorbed / mmol g

(c)

90 UiO-66(Zr)-(OH)2

IAST CO2/N2 selectivity

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

Page 14 of 26

UiO-66(Zr)-(OAc)2

75

UiO-66(Zr)-(OPr)2

60

45

30 0

20

40

60

80

100

Absolute Pressure / kPa

Figure 4. (a) CO2 sorption isotherms of UiO-66(Zr)-(OH)2, UiO-66(Zr)-(OAc)2, and UiO66(Zr)-(OPr)2 at 298 K (solid symbols, adsorption; open symbols, desorption). (b) Qst values of CO2 in the MOFs. (c) IAST selectivities of a 15/85 CO2/N2 gas mixture in the MOFs at 298 K.

ACS Paragon Plus Environment

Page 15 of 26

Stability Tests of MOFs. Excellent material robustness is expected for MOFs to be used in industrial applications under harsh conditions.7 In this study, the UiO-66(Zr)-(OPr)2 was selected as the representative esterified MOF and subjected to a series of stability tests including 1-day soaking in boiling water (100 °C), 70 °C water, highly acidic (pH = 1), and basic (pH = 11) aqueous solutions. UiO-66(Zr)-(OPr)2 demonstrates excellent water stability comparable to its MOF precursor, which is reflected in the unchanged PXRD patterns and gas sorption isotherms (Figure 5a-c).38 Nevertheless, it should be noted that some of the ester bonds in the modified MOFs were hydrolyzed after soaking in boiling water as indicated by the reduced band intensity at 1740 cm-1 in FTIR spectrum (Figure 5d). Reducing temperature to 70 °C helps to preserve the ester bonds, suggesting mild activation conditions to be adopted in further studies. (a)

(b) 250 200

Vads / cm3stp g-1

Intensity / a.u.

100ºC water treated for 1 day 70°C water treated for 1 day pH = 11 treated for 1 day pH=1 treated for 1 day UiO-66(Zr)-(OPr)2

150 100 UiO-66(Zr)-(OPr)2 pH =1 treated for 1 day pH = 11 treated for 1 day 70 °C water treated for 1 day

50 0

5

10

15

20

25

30

35

0.0

40

(c)

0.2

0.4

0.6

0.8

1.0

Relative Pressure P/P0

2 Theta / degree

(d)

3.0

2.0 1.5 1.0 UiO-66(Zr)-(OPr)2 pH = 1 treated for 1 day pH = 11 treated for 1 day 70 °C water treated for 1 day

0.5

Transmittance / a.u.

-1

2.5

Gas Uptake / mmol g

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

ACS Sustainable Chemistry & Engineering

UiO-66(Zr)-(OPr)2 pH = 1 treated for 1 day pH = 11 treated for 1 day 70°C water treated for 1 day boiling water treated for 1 day

0.0 0

20

40

60

80

Absolute Pressure / kPa

100

2200

2000

1800

1600

1400

1200

1000

Wavenumber / cm-1

Figure 5. Evaluation of the hydrothermal and chemical stability of UiO-66(Zr)-(OPr)2: (a) PXRD patterns, (b) N2 sorption isotherms at 77 K, (c) CO2 sorption isotherms at 298 K, and (d) FTIR spectra.

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

Dry Gas Breakthrough Experiments. Based on the comprehensive consideration of water affinity, CO2 uptake, and chemical stability, UiO-66(Zr)-(OPr)2 was selected as the representative esterified adsorbent for further evaluation of its postcombustion CO2 capture performance by dynamic breakthrough experiment using simulated dry flue gas. As shown in Figure S8, the MOF precursor UiO-66(Zr)-(OH)2 exhibits similar breakthrough curves using dry simulated flue gas in five consecutive breakthrough runs, indicating high material quality and good consistency with our previous results.38 The average CO2 uptake amount in these runs is 1.83 mmol g-1 (Table S1), which is slightly lower than the value obtained from single component gas sorption experiment. This trend is reasonable considering the competitive adsorption of N2 during breakthrough conditions. The CO2/N2 co-adsorption selectivity was further determined to be 29 based on the uptakes of CO2 and N2 calculated by Equation S15, which is comparable to the previously reported value (~40 depending on the gas pressure).53 Similarly, the packed bed containing UiO-66(Zr)-(OPr)2 powder exhibited decent recyclability of up to eight cycles with mild activation process in between (purging He gas through the column at room temperature for 3 hours), showing an average CO2 uptake capacity of 1.25 mmol g-1 (Figure 6, Table S1). Although the specific CO2 uptake capacity of UiO-66(Zr)-(OPr)2 is about 32% smaller than that of UiO-66(Zr)-(OH)2, the CO2 mean residence time in the column containing UiO-66(Zr)-(OPr)2 is actually only 14% lower than that of the column containing UiO-66(Zr)-(OH)2 (Table S1), suggesting that the esterification step has only a small effect on the volumetric CO2 uptake capacity which is a more important criterion for process design of postcombustion CO2 capture.7

ACS Paragon Plus Environment

Page 16 of 26

Page 17 of 26

(a) 1.0 0.8

F / F0

0.6 0.4

N2, 1st cycle CO2, 1st cycle

0.2

N2, 8th cycle CO2, 8th cycle

0.0 0

500

1000

1500

2000

2500

Time / s 1.8

60 CO2 capacity

1600

1.6

CO2/N2 adsorption selectivity

Corrected CO2 mean residence time

1200

800

400

Gas Uptake / mmol g-1

1.4 1.2

40

1.0 0.8 0.6

20

0.4 0.2

0

0.0

CO2/N2 adsorption selectivity

(b) CO2 Mean Residence Time / s

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

ACS Sustainable Chemistry & Engineering

0 1

2

3

4

5

6

7

8

Cycle

Figure 6. (a) Breakthrough curves obtained by passing a 15/85 CO2/N2 mixture gas with a flow rate of 5 sccm through a column (length: 7 cm; diameter: 0.46 cm) packed with UiO-66(Zr)(OPr)2 at 298 K with an equilibrium pressure of 1.11 bar. (b) Cyclic dry gas breakthrough performance of the above column at 298 K with an equilibrium pressure of 1.11 bar. Wet Gas Breakthrough Experiments. A straightforward strategy to exclude moisture effect on CO2 capture is to remove water in flue gas before introducing it into adsorption columns.54 The cost and energy consumed by drying flue gas can be waived if CO2 capture performance of the adsorbents is unaffected in wet conditions. Previously, we have reported that the CO2 uptake capacity of UiO-66(Zr)-(OH)2 decreased by 85% after it was pre-saturated by water vapor at 80% RH.38 However, the experimental condition of the previous study is different from practical pressure swing adsorption (PSA) processes where multiple gas components (e.g., N2, CO2 and

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

H2O) in the frontal loading compete with each other for the limited sorption sites of the unsaturated adsorbents.55-56 In this study, we conducted simulated wet flue gas breakthrough experiments without any pre-treatment of the column to mimic the industrial PSA processes, and studied how moisture in the gas stream would affect CO2 capture performance of the MOF-filled column. Inspiringly, UiO-66(Zr)-(OH)2 exhibits a CO2 uptake capacity of 1.67 mmol g-1 after two consecutive breakthrough runs in the presence of moisture with mild activation in between (Figure 7a and 7b, Table S2). The CO2 uptake capacity of MOF-filled column can be easily restored to 1.78 mmol g-1 after overnight purging of He gas at 60 °C. These values are slightly lower than that obtained from the single-component CO2 sorption isotherm (1.98 mmol g-1) or that predicted by IAST (1.96 mmol g-1). The deviation from the IAST prediction could be rationalized by the competitive adsorption of N2 and H2O together with the adsorbate-adsorbate interaction between CO2 and H2O.57 As for UiO-66(Zr)-(OPr)2, it shows a CO2 separation performance similar to that of UiO-66(Zr)-(OH)2 with CO2 uptake capacity decreasing from 1.13 mmol g-1 to 0.94 mmol g-1 in three consecutive breakthrough runs, slightly lower than the values obtained by gas sorption measurement (1.25 mmol g-1) and IAST calculation (1.20 mmol g-1). The capacity can be restored to 1.23 mmol g-1 after overnight activation (Figure 7c and 7d, Table S2). Although the CO2 uptake capacity of UiO-66(Zr)-(OPr)2 is still lower than that of UiO66(Zr)-(OH)2, it is interesting to find that the average CO2/N2 co-adsorption selectivity of UiO66(Zr)-(OPr)2 for simulated wet flue gas becomes as high as 76.6, which is more than two times of that of UiO-66(Zr)-(OH)2 (Figure 7b and 7d, Table S2). Although the propionyl groups in UiO-66(Zr)-(OPr)2 are expected to alleviate the competitive adsorption of water against CO2, we could not detect any water signal at the exit of the column during breakthrough tests, suggesting that almost all the water molecules remain trapped within UiO-66(Zr)-(OPr)2. This could be

ACS Paragon Plus Environment

Page 18 of 26

Page 19 of 26

because propionyl groups are not long enough to block the pore entrance of the MOF so that water molecules could still access the hydrophilic metal clusters and get adsorbed. In the extreme scenario where the packed column was pre-saturated with 60% RH 15/85 He/N2 mixture, UiO66(Zr)-(OPr)2 only retained a minimal CO2 uptake of 0.04 mmol g-1 (Figure S9), suggesting that

0.6 0.4 1st cycle 2nd cycle 3rd cycle 4th cycle

0.2 0.0 0

500

1000

1500

2000

2.2

1800

2.0

1600

1.8

1400 1200 1000 800 600

120 CO2 capacity CO2/N2 adsorption selectivity Corrected CO2 mean residence time 100

1.6 80

1.4 1.2

60 1.0 0.8

40

0.6

400

0.4

200

0.2

0

0.0

20

0

2500 Cycle

(d)

UiO-66(Zr)-(OPr)2 1.0 0.8 0.6 0.4 1st cycle 2nd cycle 3rd cycle 4th cycle

0.2 0.0 0

500

1000

1500

2000

1.6

1400

CO2/N2 adsorption selectivity

CO2 capacity

120

Corrected CO2 mean residence time 1200 1000 800 600 400

100

1.4

Gas Uptake / mmol g-1

(c)

CO2 Corrected Mean Residence Time / s

Time / s

1.2

80

1.0 60

0.8 0.6

40

0.4 20

200

0.2

0

0.0

2500

CO2/N2 adsorption selectivity

F / F0

0.8

2000

CO2/N2 adsorption selectivity

(b)

UiO-66(Zr)-(OH)2 1.0

Gas Uptake / mmol g-1

(a)

CO2 Corrected Mean Residence Time / s

CO2 and H2O molecules still share the same sorption sites even after postsynthetic esterification.

F / F0

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

ACS Sustainable Chemistry & Engineering

0

1

2

3

4

Cycle

Time / s

Figure 7. Breakthrough curves of the columns (length, 7 cm; diameter, 0.45 cm) packed with (a) UiO-66(Zr)-(OH)2 and (c) UiO-66(Zr)-(OPr)2 using wet 15/85 CO2/N2 mixture (60% RH) with a flow rate of 5 sccm at 298 K and a pressure drop of 0.17 and 0.24 bar across the column, respectively. Cyclic wet gas breakthrough performance of the above columns packed with (b) UiO-66(Zr)-(OH)2 and (d) UiO-66(Zr)-(OPr)2.

Conclusion

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

In summary, we adopted a scalable and facile approach combining the MHT approach and postsynthetic modification to reduce the hydrophilicity of UiO-66(Zr)-(OH)2 for postcombustion CO2 capture under wet conditions. PXRD patterns and Ar sorption isotherms reveal the intact crystallinity of the MOFs after being treated in acidic conditions during the postsynthetic esterification. Moreover, a series of stability tests show that the esterified MOF UiO-66(Zr)(OPr)2 preserves the excellent water stability of the precursor MOF UiO-66(Zr)-(OH)2. Water contact angle tests and water sorption measurements further confirm the repressed hydrophilicity of the esterified MOFs compared to UiO-66(Zr)-(OH)2. During the breakthrough experiments using simulated wet flue gas, although UiO-66(Zr)-(OPr)2 exhibits lower CO2 uptake capacity (1.08 mmol g-1) than that of UiO-66(Zr)-(OH) (1.73 mmol g-1), the CO2/N2 co-adsorption selectivity of UiO-66(Zr)-(OPr)2 (76.6) is higher than that of the latter (33.4), showing a tradeoff balance between gas uptake capacity and selectivity. Despite the continuous loss in CO2 uptake capacity during three consecutive breakthrough experiments using simulated wet flue gas, both MOFs can be easily regenerated by prolonged activation at 60 °C. It is noteworthy that esterification and the concomitant replacement of capping molecules on the Zr6 clusters cannot eliminate water adsorption in the MOFs as shown in the breakthrough experiments, suggesting that moderate hydrophobicity of MOFs cannot fully repel water molecules. Alternative esterification approaches using carbonyl chloride or anhydrides with long aliphatic chains may be promising solutions, as the aliphatic chains may not only further increase surface hydrophobicity of the MOFs but also impede water molecules from entering the internal pores of MOFs by pore channel blockings with the bulky alkyl groups. Further studies related to these approaches are undergoing in our lab to pursue water repellant adsorbents with huge potentials for practical postcombustion CO2 capture.

ACS Paragon Plus Environment

Page 20 of 26

Page 21 of 26 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

ACS Sustainable Chemistry & Engineering

Supporting Information. Materials and instrumental information, data processing methods, the breakthrough setup scheme, and additional characterization results. This material is available free of charge via the Internet at http://pubs.acs.org.

AUTHOR INFORMATION Corresponding Author *E-mail: [email protected] Notes The authors declare no competing financial interest.

ACKNOWLEDGMENT This work is supported by National University of Singapore (CENGas R-261-508-001-646), Ministry of Education - Singapore (MOE AcRF Tier 1 R-279-000-472-112, R-279-000-540114), and Agency for Science, Technology and Research (PSF R-279-000-475-305, IRG R-279000-510-305).

REFERENCES (1) Schoedel, A.; Ji, Z.; Yaghi, O. M. The role of metal–organic frameworks in a carbon-neutral energy cycle. Nat. Energy 2016, 1, 16034. (2) Oschatz, M.; Antonietti, M. A search for selectivity to enable CO2 capture with porous adsorbents. Energy Environ. Sci. 2018, 11, 57-70. (3) Reiner, D. M. Learning through a portfolio of carbon capture and storage demonstration projects. Nat. Energy 2016, 1, 15011. (4) Haszeldine, R. S. Carbon Capture and Storage: How Green Can Black Be? Science 2009, 325, 1647-1652. (5) Ho, M. T.; Allinson, G. W.; Wiley, D. E. Reducing the Cost of CO2 Capture from Flue Gases Using Pressure Swing Adsorption. Ind. Eng. Chem. Res. 2008, 47, 4883-4890.

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

(6) Ruthven, D. M.; Farooq, S.; Knaebel, K. S. Pressure Swing Adsorption; Wiley-VCH: Weinheim, 1993. (7) Pirngruber, G. D.; Llewellyn, P. L. Opportunities for MOFs in CO2 Capture from Flue Gases, Natural Gas, and Syngas by Adsorption. In Metal-Organic Frameworks: Applications from Catalysis to Gas Storage; Wiley-VCH Verlag GmbH & Co. KGaA: 2011, p 99-119. (8) Hu, Z.; Sun, Y.; Zeng, K.; Zhao, D. Structural-failure resistance of metal-organic frameworks toward multiple-cycle CO2 sorption. Chem. Commun. 2017, 53, 8653-8656. (9) Sun, Y.; Hu, Z.; Zhao, D.; Zeng, K. Mechanical Properties of Microcrystalline Metal– Organic Frameworks (MOFs) Measured by Bimodal Amplitude Modulated-Frequency Modulated Atomic Force Microscopy. ACS Appl. Mater. Interfaces 2017, 9, 32202-32210. (10) Li, J.-R.; Kuppler, R. J.; Zhou, H.-C. Selective gas adsorption and separation in metalorganic frameworks. Chem. Soc. Rev. 2009, 38, 1477-1504. (11) Yu, J.; Xie, L.-H.; Li, J.-R.; Ma, Y.; Seminario, J. M.; Balbuena, P. B. CO2 Capture and Separations Using MOFs: Computational and Experimental Studies. Chem. Rev. 2017, 117, 9674-9754. (12) Lin, Y.; Kong, C.; Zhang, Q.; Chen, L. Metal-Organic Frameworks for Carbon Dioxide Capture and Methane Storage. Adv. Energy Mater. 2017, 7, 1601296. (13) Sanz-Pérez, E. S.; Murdock, C. R.; Didas, S. A.; Jones, C. W. Direct Capture of CO2 from Ambient Air. Chem. Rev. 2016, 116, 11840-11876. (14) Wang, Y.; LeVan, M. D. Adsorption Equilibrium of Binary Mixtures of Carbon Dioxide and Water Vapor on Zeolites 5A and 13X. J. Chem. Eng. Data 2010, 55, 3189-3195. (15) 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. (16) Chanut, N.; Bourrelly, S.; Kuchta, B.; Serre, C.; Chang, J.-S.; Wright, P. A.; Llewellyn, P. L. Screening the Effect of Water Vapour on Gas Adsorption Performance: Application to CO2 Capture from Flue Gas in Metal–Organic Frameworks. ChemSusChem 2017, 10, 1543-1553. (17) Howarth, A. J.; Liu, Y.; Li, P.; Li, Z.; Wang, T. C.; Hupp, J. T.; Farha, O. K. Chemical, thermal and mechanical stabilities of metal–organic frameworks. Nat. Rev. Mater. 2016, 1, 15018. (18) O’Nolan, D.; Kumar, A.; Zaworotko, M. J. Water Vapor Sorption in Hybrid Pillared Square Grid Materials. J. Am. Chem. Soc. 2017, 139, 8508-8513. (19) McDonald, T. M.; Mason, J. A.; Kong, X.; Bloch, E. D.; Gygi, D.; Dani, A.; Crocella, V.; Giordanino, F.; Odoh, S. O.; Drisdell, W. S.; Vlaisavljevich, B.; Dzubak, A. L.; Poloni, R.; Schnell, S. K.; Planas, N.; Lee, K.; Pascal, T.; Wan, L. F.; Prendergast, D.; Neaton, J. B.; Smit, B.; Kortright, J. B.; Gagliardi, L.; Bordiga, S.; Reimer, J. A.; Long, J. R. Cooperative insertion of CO2 in diamine-appended metal-organic frameworks. Nature 2015, 519, 303-308. (20) Yan, Q.; Lin, Y.; Kong, C.; Chen, L. Remarkable CO2/CH4 selectivity and CO2 adsorption capacity exhibited by polyamine-decorated metal-organic framework adsorbents. Chem. Commun. 2013, 49, 6873-6875. (21) Hu, Y.; Verdegaal, W. M.; Yu, S.-H.; Jiang, H.-L. Alkylamine-Tethered Stable Metal– Organic Framework for CO2 Capture from Flue Gas. ChemSusChem 2014, 7, 734-737. (22) Khutia, A.; Janiak, C. Programming MIL-101Cr for selective and enhanced CO2 adsorption at low pressure by postsynthetic amine functionalization. Dalton Trans. 2014, 43, 1338-1347.

ACS Paragon Plus Environment

Page 22 of 26

Page 23 of 26 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

ACS Sustainable Chemistry & Engineering

(23) Fracaroli, A. M.; Furukawa, H.; Suzuki, M.; Dodd, M.; Okajima, S.; Gándara, F.; Reimer, J. A.; Yaghi, O. M. Metal–Organic Frameworks with Precisely Designed Interior for Carbon Dioxide Capture in the Presence of Water. J. Am. Chem. Soc. 2014, 136, 8863-8866. (24) Siegelman, R. L.; McDonald, T. M.; Gonzalez, M. I.; Martell, J. D.; Milner, P. J.; Mason, J. A.; Berger, A. H.; Bhown, A. S.; Long, J. R. Controlling Cooperative CO2 Adsorption in Diamine-Appended Mg2(dobpdc) Metal–Organic Frameworks. J. Am. Chem. Soc. 2017, 139, 10526-10538. (25) Milner, P. J.; Siegelman, R. L.; Forse, A. C.; Gonzalez, M. I.; Runčevski, T.; Martell, J. D.; Reimer, J. A.; Long, J. R. A Diaminopropane-Appended Metal–Organic Framework Enabling Efficient CO2 Capture from Coal Flue Gas via a Mixed Adsorption Mechanism. J. Am. Chem. Soc. 2017, 139, 13541-13553. (26) Liao, P.-Q.; Chen, X.-W.; Liu, S.-Y.; Li, X.-Y.; Xu, Y.-T.; Tang, M.; Rui, Z.; Ji, H.; Zhang, J.-P.; Chen, X.-M. Putting an ultrahigh concentration of amine groups into a metal-organic framework for CO2 capture at low pressures. Chem. Sci. 2016, 7, 6528-6533. (27) Moghadam, P. Z.; Ivy, J. F.; Arvapally, R. K.; dos Santos, A. M.; Pearson, J. C.; Zhang, L.; Tylianakis, E.; Ghosh, P.; Oswald, I. W. H.; Kaipa, U.; Wang, X.; Wilson, A. K.; Snurr, R. Q.; Omary, M. A. Adsorption and molecular siting of CO2, water, and other gases in the superhydrophobic, flexible pores of FMOF-1 from experiment and simulation. Chem. Sci. 2017, 8, 3989-4000. (28) Nguyen, J. G.; Cohen, S. M. Moisture-Resistant and Superhydrophobic Metal−Organic Frameworks Obtained via Postsynthetic Modification. J. Am. Chem. Soc. 2010, 132, 4560-4561. (29) Sun, Q.; He, H.; Gao, W.-Y.; Aguila, B.; Wojtas, L.; Dai, Z.; Li, J.; Chen, Y.-S.; Xiao, F.S.; Ma, S. Imparting amphiphobicity on single-crystalline porous materials. Nat. Commun. 2016, 7, 13300. (30) Sun, Y.; Sun, Q.; Huang, H.; Aguila, B.; Niu, Z.; Perman, J. A.; Ma, S. A molecular-level superhydrophobic external surface to improve the stability of metal-organic frameworks. J. Mater. Chem. A 2017, 5, 18770-18776. (31) Kitao, T.; Zhang, Y.; Kitagawa, S.; Wang, B.; Uemura, T. Hybridization of MOFs and polymers. Chem. Soc. Rev. 2017, 46, 3108-3133. (32) Zhang, W.; Hu, Y.; Ge, J.; Jiang, H.-L.; Yu, S.-H. A Facile and General Coating Approach to Moisture/Water-Resistant Metal–Organic Frameworks with Intact Porosity. J. Am. Chem. Soc. 2014, 136, 16978-16981. (33) Qian, X.; Sun, F.; Sun, J.; Wu, H.; Xiao, F.; Wu, X.; Zhu, G. Imparting surface hydrophobicity to metal-organic frameworks using a facile solution-immersion process to enhance water stability for CO2 capture. Nanoscale 2017, 9, 2003-2008. (34) Zhang, T.; Zou, B.; Shao, M.; Chen, X.; Zhang, S.; Li, L.; Du, Q.; Li, H.; Hu, Y.; Weng, J.; Xiong, W.; Zheng, B.; Zhang, W.; Huo, F. Metal-Organic Framework Wears a Protective Cover for Improved Stability. Chem. - Eur. J. 2017, 23, 7663-7666. (35) Chun, J.; Kang, S.; Park, N.; Park, E. J.; Jin, X.; Kim, K.-D.; Seo, H. O.; Lee, S. M.; Kim, H. J.; Kwon, W. H.; Park, Y.-K.; Kim, J. M.; Kim, Y. D.; Son, S. U. Metal–Organic Framework@Microporous Organic Network: Hydrophobic Adsorbents with a Crystalline Inner Porosity. J. Am. Chem. Soc. 2014, 136, 6786-6789. (36) Nguyen, N. T. T.; Furukawa, H.; Gándara, F.; Nguyen, H. T.; Cordova, K. E.; Yaghi, O. M. Selective Capture of Carbon Dioxide under Humid Conditions by Hydrophobic Chabazite-Type Zeolitic Imidazolate Frameworks. Angew. Chem., Int. Ed. 2014, 53, 10645-10648.

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

(37) Hu, Z.; Gami, A.; Wang, Y.; Zhao, D. A Triphasic Modulated Hydrothermal Approach for the Synthesis of Multivariate Metal–Organic Frameworks with Hydrophobic Moieties for Highly Efficient Moisture-Resistant CO2 Capture. Adv. Sustainable Syst. 2017, 1, 1700092. (38) Hu, Z.; Wang, Y.; Farooq, S.; Zhao, D. A Highly Stable Metal-Organic Framework with Optimum Aperture Size for CO2 Capture. AIChE J. 2017, 63, 4103-4114. (39) Yamada, T.; Kitagawa, H. Protection and Deprotection Approach for the Introduction of Functional Groups into Metal−Organic Frameworks. J. Am. Chem. Soc. 2009, 131, 6312-6313. (40) Hu, Z.; Peng, Y.; Kang, Z.; Qian, Y.; Zhao, D. A Modulated Hydrothermal (MHT) Approach for the Facile Synthesis of UiO-66-Type MOFs. Inorg. Chem. 2015, 54, 4862-4868. (41) Hu, Z.; Nalaparaju, A.; Peng, Y.; Jiang, J.; Zhao, D. Modulated Hydrothermal Synthesis of UiO-66(Hf)-Type Metal–Organic Frameworks for Optimal Carbon Dioxide Separation. Inorg. Chem. 2016, 55, 1134-1141. (42) Shearer, G. C.; Chavan, S.; Bordiga, S.; Svelle, S.; Olsbye, U.; Lillerud, K. P. Defect Engineering: Tuning the Porosity and Composition of the Metal–Organic Framework UiO-66 via Modulated Synthesis. Chem. Mater. 2016, 28, 3749-3761. (43) Stuart, B. H. Organic Molecules. In Infrared Spectroscopy: Fundamentals and Applications; John Wiley & Sons, Ltd: 2005, p 71-93. (44) Zhang, F.; Sang, X.; Tan, X.; Liu, C.; Zhang, J.; Luo, T.; Liu, L.; Han, B.; Yang, G.; Binks, B. P. Converting Metal–Organic Framework Particles from Hydrophilic to Hydrophobic by an Interfacial Assembling Route. Langmuir 2017, 33, 12427-12433. (45) Hu, Z.; Zhao, D. De facto methodologies toward the synthesis and scale-up production of UiO-66-type metal–organic frameworks and membrane materials. Dalton Trans. 2015, 44, 19018-19040. (46) Hu, Z.; Faucher, S.; Zhuo, Y.; Sun, Y.; Wang, S.; Zhao, D. Combination of Optimization and Metalated-Ligand Exchange: An Effective Approach to Functionalize UiO-66(Zr) MOFs for CO2 Separation. Chem. - Eur. J. 2015, 21, 17246-17255. (47) Rowsell, J. L. C.; Yaghi, O. M. Effects of Functionalization, Catenation, and Variation of the Metal Oxide and Organic Linking Units on the Low-Pressure Hydrogen Adsorption Properties of Metal−Organic Frameworks. J. Am. Chem. Soc. 2006, 128, 1304-1315. (48) Nugent, P.; Belmabkhout, Y.; Burd, S. D.; Cairns, A. J.; Luebke, R.; Forrest, K.; Pham, T.; Ma, S.; Space, B.; Wojtas, L.; Eddaoudi, M.; Zaworotko, M. J. Porous materials with optimal adsorption thermodynamics and kinetics for CO2 separation. Nature 2013, 495, 80-84. (49) Shekhah, O.; Belmabkhout, Y.; Chen, Z.; Guillerm, V.; Cairns, A.; Adil, K.; Eddaoudi, M. Made-to-order metal-organic frameworks for trace carbon dioxide removal and air capture. Nat. Commun. 2014, 5, 4228. (50) Wang, Y.; Zhao, D. Beyond Equilibrium: Metal–Organic Frameworks for Molecular Sieving and Kinetic Gas Separation. Cryst. Growth Des. 2017, 17, 2291-2308. (51) Hu, Z.; Zhang, K.; Zhang, M.; Guo, Z.; Jiang, J.; Zhao, D. A Combinatorial Approach towards Water-Stable Metal–Organic Frameworks for Highly Efficient Carbon Dioxide Separation. ChemSusChem 2014, 7, 5. (52) Cmarik, G. E.; Kim, M.; Cohen, S. M.; Walton, K. S. Tuning the Adsorption Properties of UiO-66 via Ligand Functionalization. Langmuir 2012, 28, 15606-15613. (53) Yang, Q.; Vaesen, S.; Ragon, F.; Wiersum, A. D.; Wu, D.; Lago, A.; Devic, T.; Martineau, C.; Taulelle, F.; Llewellyn, P. L.; Jobic, H.; Zhong, C.; Serre, C.; De Weireld, G.; Maurin, G. A Water Stable Metal–Organic Framework with Optimal Features for CO2 Capture. Angew. Chem., Int. Ed. 2013, 52, 10316-10320.

ACS Paragon Plus Environment

Page 24 of 26

Page 25 of 26 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

ACS Sustainable Chemistry & Engineering

(54) Riboldi, L.; Bolland, O.; Ngoy, J. M.; Wagner, N. Full-plant analysis of a PSA CO2 capture unit integrated in coal-fired power plants: post-and pre-combustion scenarios. Energy Procedia 2014, 63, 2289-2304. (55) Riboldi, L.; Bolland, O. Overview on pressure swing adsorption (PSA) as CO2 capture technology: state-of-the-art, limits and potentials. Energy Procedia 2017, 114, 2390-2400. (56) Li, G.; Xiao, P.; Webley, P.; Zhang, J.; Singh, R.; Marshall, M. Capture of CO2 from high humidity flue gas by vacuum swing adsorption with zeolite 13X. Adsorption 2008, 14, 415-422. (57) Myers, A. L.; Prausnitz, J. M. Thermodynamics of mixed-gas adsorption. AIChE J. 1965, 11, 121-127.

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

Synopsis Esterified zirconium metal–organic frameworks with increased hydrophobicity can enhance the CO2/N2 co-adsorption selectivity during breakthrough experiments using simulated wet flue gas, suggesting their huge potential in practical postcombustion CO2 capture.

ACS Paragon Plus Environment

Page 26 of 26