Controllable Adsorption of CO2 on Smart Adsorbents: An Interplay

2 days ago - Photoresponsive adsorbents for CO2 capture attract significant attention, because the variation of adsorption capacity can be tailored co...
0 downloads 6 Views 1MB Size
Subscriber access provided by UNIV OF NEW ENGLAND ARMIDALE

Controllable Adsorption of CO2 on Smart Adsorbents: An Interplay between Amines and Photoresponsive Molecules Lei Cheng, Yao Jiang, Shi-Chao Qi, Wei Liu, Shu-Feng Shan, Peng Tan, Xiao-Qin Liu, and Lin-Bing Sun Chem. Mater., Just Accepted Manuscript • DOI: 10.1021/acs.chemmater.8b01005 • Publication Date (Web): 24 Apr 2018 Downloaded from http://pubs.acs.org on April 24, 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 10 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

Controllable Adsorption of CO2 on Smart Adsorbents: An Interplay between Amines and Photoresponsive Molecules Lei Cheng,‡ Yao Jiang,‡ Shi-Chao Qi, Wei Liu, Shu-Feng Shan, Peng Tan, Xiao-Qin Liu, and Lin-Bing Sun* State Key Laboratory of Materials-Oriented Chemical Engineering, Jiangsu National Synergetic Innovation Center for Advanced Materials (SICAM), College of Chemical Engineering, Nanjing Tech University, Nanjing 210009, China. ABSTRACT: Photoresponsive adsorbents for CO2 capture attract significant attention, because the variation of adsorption capacity can be tailored conveniently by light irradiation. However, all photoresponsive adsorbents reported until now capture CO2 via physical interaction (weak sites). Thus the selectivity of CO2 over other gases like CH4 is low, and the variation of CO2 adsorption capacity is originated from pure structural change via isomerization of photoresponsive units. Despite great challenges, photoresponsive adsorbents for CO2 capture based on chemical interaction (strong sites) are extremely desirable. Here, we report for the first time the construction of photoresponsive adsorbents with chemical adsorption sites for CO 2. The control of adsorption capacity is based on the interplay between active amine sites {A, [3-(methylamine)propyl]trimethoxysilane} and photoresponsive azobenzene molecules [P, 4-(3-triethoxysilylpropyl-ureido)azobenzene]. Density functional theory (DFT) calculations reveal that the negative surface potential of amines correlate with the adsorption capacity on CO 2. Upon visible light (450 nm) irradiation, trans-isomers of azobenzene are formed and interacted with amines, which leads to decreased surface electrostatic potential of amines and CO2 can thus be adsorbed freely on exposed active sites. In contrast, ultraviolet light (365 nm) irradiation results in the isomerization of azobenzene from tran to cis conformation. The surface electrostatic potential of amines increases obviously and the active sites are thus blocked. A maximum variation amount is obtained on the adsorbent with comparable density of A and P (0.34 group nm‒2), which confirms the interplay between amine sites and photoresponsive molecules. Because amines are specific active sites for CO2, high selectivity of CO2 over CH4 is obtained, and such selectivity is tunable upon irradiation with UV/VIS light. We thus demonstrate the successful control of CO2 captured by chemical sites through the interplay between amines and azobenzene molecules, which is impossible or difficult to realize via conventional structural change caused by isomerization of photoresponsive units. INTRODUCTION Natural gas, a cleaner energy for industrial and residential heating, has become increasingly important because of the evergrowing global demand.1,2 The active component of natural gas is methane (CH4) while one of the main impurities is carbon dioxide (CO2).3 The existence of CO2 reduces the reactivity and heating capacity; it also causes corrosion of the relevant pipeline and equipment.4 Indeed, the purification of natural gas is a very substantial separation engineering in industrial processes, aiming to meet equipment specifications and relevant standards on caloric value.5,6 Adsorption is considered a highly promising technique for the removal of CO2 from natural gas.7 It is noticeable that conventional adsorbents possess fixed active sites, and the difference in adsorption capacity comes from the variation of pressure and/or temperature.8 The difference in adsorption capacity under different conditions is then used for the design of cyclic adsorption processes. However, the variation of pressure/temperature is usually energy-intensive and new methods to control adsorption capacity is expected.9 Much attention has been given to the functional materials responsive to external stimuli such as temperature, redox potential, pH, and light.10-23 Among diverse stimulation forms, light is considered the most attractive one because it is a rapid and remote stimulus with high precision.24-36 Moreover, irradiation with light does not generate any undesired by-

products.37-40 As a result, the photoresponsive system has been explored for potential applications in CO2 adsorption.41,42 One interesting example was reported by Park et al.;41 they synthesized a metal-organic framework (MOF) by using a photoresponsive ligand 2-(phenyldiazenyl)terephthalate. The obtained MOF realized reversible alteration of CO2 capture upon photochemical treatment. Another attractive example from Lyndon et al. reported the conversion of azobenzene to 4,4’-dicarboxylate, which was utilized to construct a MOF. 42 The MOF decorated with photoresponsive motifs could undergo dynamic photo-induced structural versatility, leading to an obvious change for the adsorption of CO2. It is worth noting that the adsorption isotherms of CO2 are almost linear in both cases.41,42 This indicates that the adsorbate-adsorbent interaction is quite weak and physical adsorption is predominant, which is not beneficial to selective capture of CO2. Actually, in all literature the change in CO2 adsorption capacity is derived from pure structural variation via isomerization of photoresponsive groups, due to the lack of suitable active sites such as amines. From the viewpoint of practical applications, photoresponsive adsorbents for CO2 capture based on strong adsorbate-adsorbent interaction (e.g. chemical interaction) are extremely desirable. This may result in controllable adsorption of CO2 with high selectivity from gas mixtures. Noteworthy, it is difficult to adjust the adsorption of CO2 that interacted strongly with active sites through traditional structural variation induced by photoresponsive molecules. It is interesting to note

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

that Lee et al. co-functionalized amine with azobenzene groups onto the gold nanoparticles.43 This renders the nanomaterials responsive to light irradiation and CO2 and the reversible assembly-disassembly behaviors, while the adsorption of CO2 was not mentioned. Apparently, the development of photoresponsive adsorbents for CO2 capture based on chemical adsorption is a great challenge. Here, we report the fabrication of photoresponsive adsorbents with chemical adsorption sites for CO2, for the first time. The adjustment of adsorption capacity is originated from the interplay between active sites (amine) and photoresponsive molecules (azobenzene). This is obviously different from pure structural variation via isomerization of photoresponsive molecules for the adsorbents reported in literature.41,42 The adsorbents were constructed by incorporating amines {A, [3(methylamine)propyl]trimethoxysilane} and photoresponsive molecules [P, 4-(3-triethoxysilylpropyl-ureido)azobenzene, Figure S1] simultaneously into mesoporous silica (MS) as shown in Scheme 1. It is known that azobenzene is a photoresponsive molecule with short response time and reversibility between trans/cis conformation.44 The trans and

Page 2 of 10

cis isomerization of the azo bond is induced by corresponding irradiation with visible (VIS) and ultraviolet (UV) light. Our calculations from density functional theory (DFT) indicate that the negative surface potential of amines correlate with the adsorption capacity on CO2. Upon VIS light (450 nm) irradiation, azobenzene molecules isomerize from cis to trans conformation, which leads to decreased surface electrostatic potential of amines and CO2 can be adsorbed freely on exposed active sites. Consequent UV light (365 nm) irradiation converts azobenzene molecules from trans to cis conformation. The surface electrostatic potential increases obviously and the active sites are thus blocked. Apparently, the interplay between active sites and photoresponsive molecules causes the change of surface electrostatic potential of amines, and subsequently tailors the adsorption behavior of CO2. Because the amines are specified active sites for CO2 rather than CH4, high selectivity is thus obtained on the present photoresponsive adsorbents. This is the first report of photoresponsive adsorbents that can realize the control of CO2 adsorption on amines through chemical interaction as well as high selectivity of CO2/CH4.

Scheme 1. Control of CO2 Adsorption Behavior on the Smart Adsorbents through An Interplay Between Amines (A) and Photoresponsive Molecules (P). (A) VIS Light Irradiation Leads to Trans Isomerization of Azobenzene Molecule, Decreased Surface Electrostatic Potential, and Exposure of Active Sites; (B) UV Light Irradiation Leads to Cis Isomerization of Azobenzene Molecule, Increased Surface Electrostatic Potential, and Blockage of Active Sites

EXPERIMENTAL SECTION Chemicals. Toluene (≥99.5%) and tetrahydrofuran (99%) were purchased from Sinopharm and Adamas-beta respectively, and were dehydrated before use. Methanol (≥99.9%), and hexane (Aladdin, 97%) were purchased from Aladdin, and were used as received. 4-aminoazobenzene (98%) was purchased from TCI Co. Ltd. Cetyltrimethylammonium bromide (CTAB; 99%), 3-(triethoxysilyl)propyl isocyanate (95%), [3(methylamine)propyl]trimethoxysilane (95%), and tetraethylorthosilicate (98%) were purchased from SigmaAldrich Reagent Co. Ltd. Deionized water was used for all of

the experiments. Materials Sythesis. The photoresponsive molecule (P) precursor 4-(3-triethoxysilylpropyl-ureido)azobenzene was synthesized according to literature procedure.45 In a typical process, 3-(triethoxysilyl)propyl isocyanate (2.05 g, 8.12 mmol) and 4-aminoazobenzene (1.58 g, 7.85 mmol) was dissolved in tetrahydrofuran (12 mL) and reflux under N2 for 24 h. Then, hexane was added to the solution at 20 °C. After filtration, and being washed with hexane several times, the precipitate was dried at 60 oC for 24 h under vacuum. The MS was prepared according to a previously described

2 ACS Paragon Plus Environment

Page 3 of 10 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

method.46 A mixture of CTAB (1.0 g, 2.74 mmol), NaOH (2 M, 3.5 mL), and 480 mL deionized water was heated to 80 °C, and tetraethylorthosilicate (5.0 mL, 23.0 mmol) was added dropwise to the solution. The mixture was then stirred at 80 °C for 2 h. After filtering, and washing with deionized water and methanol, the as-synthesized MS was obtained. In order to remove the surfactant template (CTAB) from as-synthesized MS, the sample was suspended in a mixture of methanol (100 mL) and concentrated HCl (1.0 mL, 2M), and the solution was heated under reflux for 24 h. The precipitate was then filtered and washed extensively with H2O and methanol and placed under vacuum to remove the remaining solvents to yield MS. Smart adsorbents containing methylamine group and azobenzene units were prepared as follows. The MS (0.5 g) was dispersed in 80 mL toluene with ultrasonic for 5 min. Then, the A precursor [3-(methylamine)propyl]trimethoxysilane (70, 85, or 100 mg) and the P precursor 4-(3-triethoxysilylpropylureido)azobenzene (150, 210, or 370 mg) were added to the mixture. Before the reaction, the obtained solution was soaked for 24 h to make adequate contact with each other. After the mixture was stirred at 80 °C under N2 for 12 h, the solid was filtrated and washed with methanol several times and dried under vacuum for 24 h. The resultant samples were denoted as AmPn@MS samples, where m varies from 1 to 3 and corresponds to the A content ranged from 0.26 to 0.42 group nm2, and n varies from 1 to 3 and corresponds to the P content ranged from 0.25 to 0.67 group nm2. The grafted amount of A and P was calculated according to elemental content of nitrogen and TG analysis. Afterwards, the density of A or P can be obtained by combining the grafted amount of A or P and the surface areas of AmPn@MS. Materials Characterization. 1H NMR spectra of samples were recorded on a 400 MHz nuclear magnetic resonance instrument (Bruker Avance). UV-vis diffuser reflectance spectra (UV-vis DRS) were collected on the Shimadzu UV devices. X-ray diffraction patterns of samples were recorded on Rigaku D/MAX-γA. Transmission electron microscopy and EDX analyses were performed using an FEI Tecnai G2 F20 electron microscope operated at 200 kV. Scanning transmission electron microscopy images were collected on Hitachi SU8010 equipment. Nitrogen sorption isotherms were obtained from ASAP 2020 analyzer at 77 K. The samples were degassed at 80 °C for 24 h prior to analysis. The Brunauer-Emmett-Teller (BET) surface area was calculated using adsorption data in a relative pressure ranging from 0.04 to 0.20. The total pore volume was determined from the amount adsorbed at a relative pressure of 0.99. The pore size distributions were calculated from the adsorption isotherms using the Barrett-Joyner-Halenda (BJH) methods. Liquid UV-vis spectrophotometer spectra were collected on the PerkinElmer Lambda 35 series devices. Fourier transform infrared spectra (IR) were carried out on a Nicolet Nexus 470 spectrometer. The C, H, and N elements analysis were performed by a Vario Micro Cube elemental analyzer. Thermogravimetric analysis was performed in a flow of N2 (20 mL min−1) from room temperature to 1000 °C on a NETZSCH thermobalance (STA-499C F3). Adsorption Tests. Adsorption of CO2 and CH4 was conducted on ASAP 2020 analyzer. CO2 (99.999%) and CH4 (99.999%) gases were used for the adsorption measurements. The free space was determined using He (99.999%), assuming

that He could not be adsorbed at the temperatures investigated. The gas adsorption experiments were measured using ASAP 2020 analyzer. After degas, the samples (in the quartz tube) were irradiated with UV light (wavelength 365 nm) or VIS light (wavelength 450 nm) for about 3 h by using a xenon lamp (CEL-HXUV300) equipped with a filter, and then used for the adsorption measurement directly. The adsorption isotherms of CO2 and CH4 at 0 °C and 25 °C were measured in a water bath with circulating water equipment. The isosteric heats of CO2 adsorption (Qst) were calculated from the CO2 adsorption isotherms at temperatures of 0 oC and 25 oC, the data were simulated with Virial expression composed of parameters ai and bi that are independent of temperature in terms of equation 1. In general, a nonlinear curve was obtained displaying the connection between lnP and adsorption quantity (N), from the fitting parameters results of ai, the Qst was calculated in terms of equation 2.47 In the equations, P is pressure, N is amount adsorbed, T is temperature, and m and n represent the number of parameters a and b (where m ≤ 5 and n ≤ 2). 1

ln P = ln N + ∑mi=0 Ni + ∑ni=0 bi Ni

(1)

Qst = − R ∑mi=0 ai Ni

(2)

T

To investigate the adsorption selectivity of CO2 over CH4 on adsorbents, the selectivity is defined as S = (x1/y1)/(x2/y2), where x and y represent the composition of the adsorbed phase and the gas phase, respectively. The ideal adsorption solution theory (IAST) of Myers has been reported to predict binary gas mixture adsorption in porous materials accurately, 48 and the dual-site Langmuir (DL) model was chosen to fit the adsorption isotherms, and then DL-IAST was utilized to estimate CO2/CH4 selectivity of adsorbents. RESULTS Synthesis and Characterization. The photoresponsive unit (P) precursor was first synthesized, and the structure was confirmed by nuclear magnetic resonance (NMR, Figure S2). The precursors of P and A were then incorporated via silanol groups in MS. Corresponding density of A and P is also calculated based on their content and surface area of the support (Table 1). Structural characterization of A and P-functionalized adsorbents was performed using various methods. The lowangle X-ray diffraction (XRD) patterns of samples are presented in Figure 1. It can be found that the hexagonal pore structure of MS (p6mm) is well preserved after modification. With increasing amount of organic units, the intensity of the reflections become low gradually. In the wide-angle XRD patterns (Figure S3), a single broad diffraction peak centred at 23o is detected due to the amorphous silica walls. The incorporation of P and A does not bring new diffraction peaks, suggesting the lack of long-range order. In N2 adsorptiondesorption isotherms, all samples display typical type-IV isotherms, corresponding to mesoporous structure (Figures S4S6). Figures S5 and S6 exhibit that the pore size of AmPn@MS decreases due to the introduction of organic units. As listed in Table 1, the surface areas and pore volumes decline progressively with the increase of A and P loading, while open channels and high surface areas are still maintained.

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

Table 1. Textual properties and composition of different samples Elemental composition (wt%) C

H

N

Density of A (group nm2)

1.46

2.01

1.68

0.04

0

0

837

0.86

10.47

1.95

3.07

0.26

0.33

A2P2@MS

797

0.78

10.83

2.01

3.18

0.34

0.34

A3P2@MS

749

0.74

11.78

2.18

3.42

0.42

0.33

A2P1@MS

815

0.88

10.28

1.91

3.01

0.34

0.25

A2P3@MS

768

0.76

12.63

2.21

3.69

0.35

0.67

SBET

Vp

(m2 g1)

(cm3 g1)

MS

1365

A1P2@MS

Sample

Density of P (group nm2)

Electron microscopy was utilized to reveal the morphology and pore structure of the adsorbents. As shown in scanning electron microscopy (SEM) images (Figure S7), the samples give spherical morphologies before and after modification. In transmission electron microscopy (TEM) images, the pristine MS exhibits the regular pore periodicity and straight pore channels (Figure 2A). After modification, the periodic ordering of mesostructure is preserved to some extent (Figure 2B). The presence of C, N, O, and Si was examined in the energy dispersive X-ray (EDX) spectrum for A2P2@MS (Figure S8). Moreover, a homogeneous distribution of C, N, O, and Si is evidenced in the elemental mapping image (Figure 2C). According to these results, it is certain that the prepared composites are comparable to pristine MS in the matter of morphology and mesostructure.

Figure 1. Low-angle XRD patterns of the samples before and after loading different amounts of A and P.

To confirm the successful introduction of A and P molecules, infrared (IR) spectra of different samples were collected (Figure 3). Several new vibrational bands are detected in addition to the typical bands of silica. The vibrational bands at 1480 cm −1 is ascribed to –NH– stretching modes of the A moieties. The bands at 1530 and 690 cm−1 are derived from the –NH–CO– NH– stretching vibration and the skeletal vibrations of the P moieties, respectively.49 The intensity of such peaks increases gradually with increasing content of A (from A1P2@MS to A2P2@MS and A3P2@MS) and P (from A2P1@MS to A2P2@MS and A2P3@MS). TG analysis of the adsorbents is depicted in Figure S9. The weight loss at temperatures up to

Page 4 of 10

100 °C belongs to the removal of physically adsorbed water in the samples. From 125 to 300 °C, the weight loss attributed to the dehydration of silanol condensation is detected. Subsequently, the A and P decompose gradually from 300 to 700 °C. Elemental analysis indicates that the content of C, H, and N elements originated from A and P increases with the growth of their loading (Table 1).

Figure 2. TEM images of the samples (A) MS and (B) A2P2@MS as well as (C) EDX-mapping images of C, N, O, and Si elements.

On the basis of the above-mentioned results, it is apparent that both the active sites (A) and photoresponsive molecules (P) are successfully incorporated into MS. Moreover, the morphology and ordered mesostructure are well-maintained regardless of the loading amounts of P and A in the adsorbents.

Figure 3. IR spectra of the samples before and after loading different amounts of A and P.

Photoresponsive Properties. The photoresponsive property of the azobenzene derivative (P) was first evaluated by using UV-vis spectrometer. Figure S10 shows the absorption spectra of azobenzene derivative exposed to UV or VIS light for different time. The absorption bands of azobenzene derivative appears at around 450 and 350 nm, which belongs to the nπ* transition of the cis isomer and ππ* transition of the trans isomer, respectively.50 With the prolonging time for UV light irradiation, the band at 350 nm decreases apparently and the band at 450 nm increases, which indicates a progressive

4 ACS Paragon Plus Environment

Page 5 of 10 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

transformation from trans to cis isomers.51 A photostationary state is reached after 2 min, and it appears no further change upon irradiation for longer time. Upon VIS light irradiation, the absorption band at 350 nm restores the initial state (Figure S11). Remarkably, the band intensity even surpasses that of the initial sample slightly, since there are a fraction of cis-isomers in the initial sample. Moreover, the photoisomerization change of the azobenzene derivative is reversible through alternating UV and VIS light irradiation.

irradiation, a new band emerges at 699 cm1 because of the cis isomerization.14,52 Afterwards, the trans isomer can be restored after VIS light irradiation, accompanied by the disappearance of the band at 699 cm1.

Figure 5. Adsorption isotherms of CO2 and CH4 on A2P2@MS at 0 oC upon VIS and UV light irradiation.

Figure 4. (A) Changes in the UV-vis spectra of A2P2@MS upon irradiation with UV and VIS light. (B) Reversible changes in absorbance at 370 nm as a function of cycles for A2P2@MS.

Afterwards, the photoresponsive properties of A and Pfunctionalized adsorbents were evaluated. For the P units in trans conformation, a signature adsorption band at 370 nm is observed in the UV-vis spectra (Figures 4A and S12). After 3 min UV light irradiation, the absorption band in the 370 nm decreases, and the sample reaches a photostationary state (cis conformation) after about 5 min UV light irradiation. Inversely, through VIS light irradiation, the trans conformation is completely restored (Figure S13). In addition, the efficient trans/cis isomerization of photoreponsive units in the A and Pfunctionalized adsorbents is totally reversible upon UV and VIS light irradiation (Figures 4B). The XRD and IR data upon UV irradiation and visible irradiation were collected. The XRD patterns of the sample upon UV/VIS irradiation are almost the same as shown in Figure S14, indicating that the mesoporous structure of MS are not affected by isomerization of azobenzene groups. Interestingly, the trans/cis isomerization of the composites can be refleced by IR spectra (Figure S15). The IR spectrum of A2P2@MS shows a band at 690 cm1, which is attributable to the trans isomer of azobenzene. After UV light

Figure 6. Changes of CO2 adsorption amount upon VIS and UV light irradiation on adsorbents with different loading of (A) P and (B) A. The numerals above the bars represent the density of P (in Figure A) and A (in Figure B) in the adsorbents with an unit of group nm‒2.

The aforesaid results demonstrate that the photoresponsive properties of azobenzene molecules are well-retained after incorporating into the adsorbents together with amines. The A

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

and P-functionalized adsorbents exhibit excellent reversibility in trans/cis isomerization upon UV and VIS light irradiation. Adsorption Behavior. The adsorption behavior of CO2 and CH4 on different adsorbents upon UV (365 nm) and VIS light (450 nm) irradiation was systematically studied. For comparison, two adsorbents functionalized separately with A and P were prepared. The content of A or P in the two adsorbents is identical to that in A2P2@MS, and thus the adsorbents are denoted as A2@MS and P2@MS respectively (corresponding characterizatoin results are shown in Figure S16). The adsorption of CO2 on pristine MS, A2@MS and P2@MS were first examined (Figures S17-S19). Because of the absence of active sites for chemical adsorption, pristine MS gives an almost linear adsorption isotherm. Additionally, the change of adsorption amount upon UV and VIS light irradiation is negligible. The loading of amines presents chemical adsorption sites, and thus obvious adsorption of CO2 at low pressure is observed in A2@MS. However, the sample A2@MS exhibits unchanged adsorption amount of CO 2 during UV and VIS light irradiation, since it does not have any photoresponsive groups. There are photoresponsive units but no amines in the sample P2@MS, thus UV and VIS light irradiation cannot cause any change in CO2 adsorption as well. Interestingly, A and Pfunctionalized adsorbents exhibit quite different adsorption behavior upon light irradiation. Taking A2P2@MS as an example, the adsorption amount of CO2 is 72.4 mg g−1 upon UV light irradiation while increases to 112.3 mg g−1 upon VIS light irradiation (Figure 5). This corresponds to 35.6% of change for the adsorption amount of CO2. In addition, the adsorption of CH4 on the A2P2@MS was examined. In general, the adsorption amount of CH4 is low, due to the functionalization of specific active sites (amines) for CO2 and the absence of active sites for CH4. Moreover, there is almost no detectable change of CH4 adsorption upon UV and VIS light irradiation, suggesting that the isomerization of photoresponsive units has no effect on the adsorption of CH4. The adsorption of CO2 on a series of A and P-functionalized adsorbents with different density was studied (Figures 6 and S20-S24). Among the adsorbents studied, the change amount of CO2 adsorption reaches a maximum when the density of A and P is comparable (namely, the density is 0.34 group nm‒2 for both units).

Page 6 of 10

Figure 7. After the fresh adsorbent was saturated with CO2, regeneration was conducted. The regenerated adsorbent was irradiated with UV/VIS light, and then reused for adsorption. The adsorption amount of CO2 in four cycles was recorded. It is noticeable that the adsorption amount as well as the change between UV and VIS light irradiation almost keep at the same level. This demonstrates the high reversibility of the present photoresponsive adsorbents during UV and VIS light irradiation. The IAST model was employed to estimate the selectivity of CO2/CH4 as displayed in Figure 8A. Upon VIS light irradiation, the CO2/CH4 selectivity on A2P2@MS at 0 °C and 0.1 bar is 11293. Upon UV light irradiation, the CO2/CH4 selectivity declines to 260 under the same conditions. The results suggest high selectivity of CO2 over CH4. Futhermore, such selectivity is tunable upon irradiation with UV and VIS light irradiation. In addition, as shown in Table S1, A2P2@MS exhibited excellent selectivity toward CO2 with CH4, in comparison to other amino-functionalized adsorbents. To further understand the affinity of adsorbents with CO2, the isosteric heat of adsorption was calculated (Figure 8B). UV/VIS light irradiation causes the difference in heat of adsorption obviously. Upon VIS light irradiation, the heat of adsorption is 59.4 kJ·mol−1, which decreases to 48.9 kJ·mol−1 upon UV irradiation.

Figure 8. (A) IAST selectivity of CO2/CH4 on A2P2@MS and (B) CO2 isosteric heat of adsorption of the A2P2@MS upon UV and VIS light irradiation. Figure 7. Adsorption cycles of CO2 over A2P2@MS upon UV and VIS light irradiation.

The reversibility of the CO2 adsorption by alternating irradiation with UV and VIS light was evaluated as shown in

The foregoing results indicate that the adsorption of CO 2 is tunable on A and P-functionalized adsorbents upon UV/VIS irradiation, which is impossible to realize for the adsorbents functionalized with individual A or P. The interplay between A and P is confirmed by that a maximum change of CO2

6 ACS Paragon Plus Environment

Page 7 of 10 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 is obtained at comparable density of A and P. Because the active sites (A) incorporated is specific for CO2, CH4 is seldom adsorbed. This leads to high selectivity of CO2/CH4, which can be tailored by UV/VIS irradiation as well. Furthermore, the different adsorption behavior on A and Pfunctionalized adsorbents is totally reversible upon irradiation with corresponding UV or VIS light. DISCUSSION Albeit important, the development of photoresponsive adsorbents for CO2 capture based on strong adsorbate-adsorbent interaction has remained a great challenge. In this work, we report the fabrication of photoresponsive adsorbents with chemical adsorption sites for CO2 by incorporating chemical adsorption sites (amines) and azobenzene molecules simultaneously. The adsorption of CO2 on A and Pfunctionalized adsorbents is controllable upon UV/VIS light irradiation. The mechanism for the adjustment of CO 2

adsorption is explored. The steric effect caused by isomerization of azobenzene molecules is first considered.41 Upon VIS light irradiation, trans-azobenzene is formed and seldom shows steric hindrance of amines. The exposed active sites can thus capture CO2. Upon UV light irradiation, the transazobenzene molecules are converted to their cis isomers. The enhanced steric hindrance is possible to block some active sites. As a result, the adsorption amount of CO2 diminishes. Nevertheless, the pore size of the MS support is about 3 nm, and the size of benzene ring in azobenzene is 0.6 nm. Perhaps the azobenzene derivatives block the pores of MS partially, but CO2 molecule has a kinetic diameter of 0.33 nm can also enter the pores freely. In addition, the benzene ring with planar structure is impossible to fully shield amine sites and subsequently obstruct the adsorption of CO2. Therefore, the steric effect caused by isomerization of azobenzene should play a minor (if any) role on the control of CO2 adsorption.

Figure 9. Simulated geometries and surface electrostatic potentials of (A) methylamine group, (B) trans-azobenzene interacted with methylamine group, and (C) cis-azobenzene interacted with methylamine group.

Based on the above analysis, an interplay between amines and azobenzene molecules is believed to exist in the present system. DFT calculations were carried out to disclose the mechanism. The fully relaxed molecular geometries and the surface electrostatic potentials are calculated. With Grimme’s dispersion correction of D3 version, Becke’88 exchange and Lee-Yang-Parr correlation functional at Def2-SVP basis sets level are employed. The convergence criteria of maximum force and root mean square force are set as 1.5 × 10−5 and 1.0 × 10−5 a.u., respectively. The simulated geometries and surface electrostatic potentials of azobenzene and methylamine groups were shown in Figure 9. The calculated free energy of formation of trans-azobenzene-methylamine and cis-azobenzenemethylamine are 8.6 and 6.3 kcal mol1, respectively, suggesting that both trans- and cis-azobenzene can well interact with methylamine groups. From the perspective of thermodynamics, both trans- and cis-azobenzene groups are inclined to interact with methylamine groups by van der Waals force. For trans-azobenzene-methylamine, the azo atoms and the hydrogen atoms of secondary amino are the main sites for the van der Waals interaction (Figure 9B). But for cis-

azobenzene-methylamine, the main sites are the hydrogen atom of benzene ring 1 in azobenzene group and the hydrogen atom of methylamine groups (Figure 9C). Normally, the surface electrostatic potentials can be used to analyze the active sites of CO2 adsorption. As shown in Figure 9A, the negative surface potential of methylamine groups reaches 0.042 eV, which can be regarded as the main region for CO2 adsorption. After VIS light irradiation, the diminished active sites are negligible because of the electrostatic potential near the methylamine groups is still around 0.04 eV. Inversely, the benzene ring 2 has a strong chemical shielding effect on methylamine groups after UV irradiation, and the electrostatic potential near the methylamine group suddenly increases to 0.01 eV and even to electrically neutral, which leads to the substantial loss of active sites for CO2 adsorption. Obviously, the controllable CO2 adsorption on A and Pfunctionalized adsorbents can be attributed to the interplay between amines and azobenzene molecules. Upon VIS irradiation, trans-isomers of azobenezene are formed and interacted with amines, while the negative electrostatic

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

potential of amines is still comparable to free amines. Hence, a high adsorption amount of CO2 is obtained. Irradiation with UV light results in the formation of cis-azobenzene, and the surface electrostatic potential of amines increases sharply due to the interaction with cis-azobenzene. Thus, the adsorption amount of CO2 declines. It is worthy of note that the reference adsorbents functionalized with individual A or P do not show any change on CO2 adsorption upon UV/VIS irradiation. This indicates the significance of both A and P in the control of CO2 adsorption. The effect of density of A and P on the control of adsorption behavior was examined as well. It is demonstrated that the A2P2@MS with comparable density of A and P gives the best performance. This thus confirms the interplay between amines and azobenzene molecules. Because amines are specific active sites for CO2, the adsorption amount of CH4 on the same adsorbents is quite low. Moreover, the change amount of CH 4 upon irradiation with UV and VIS is neglectable. This leads to high CO2/CH4 selectivity, and the selectivity over the adsorbents upon VIS irradiation is evidently higher than that upon UV irradiation. Similarly, the enhanced surface electrostatic potential of amines leads to decreased interaction between amines and CO2, and thus the heat of adsorption upon UV irradiation is lower than that upon VIS irradiation. In addition, another possible mechanism can be discussed. There is a hydrogen-bond interaction between the azobenzene groups and the amines.53 The cis-azobenzene groups were shown to exhibit strong affinity towards hydrogen-bond donors. Methylamine, a secondary amine, which can be act as a hydrogen-bond donor. Hence, a low adsorption amount of CO 2 is obtained because of no affinity to amines. In contrast, the trans-azobenzene groups had no appreciable affinity to amine groups, and the isolated amine groups can be combined with the CO2 molecules freely, resulting in an enhanced adsorption amount of CO2. In short, the controllable adsorption of CO2 on A and Pfunctionalized adsorbents is ascribed to the interplay between A and P. Upon irradiation with UV and VIS light, the mode for the interaction between A and P changes, which can be reflected quantitatively by the surface electrostatic potential through DFT calcualtions. The steric effect of azobenzene molecules on the adsorption of CO2 should be tiny, if any. CONCLUSION We have demonstrated the successful fabrication of photoresponsive adsorbents with chemical adsorption sites for CO2 for the first time. Photoresponsive azobenezene molecules (P) and active amine sites (A) are incorporated into mesoporous silica simultaneously. Density functional theory calculations indicate that the interplay between active sites and photoresponsive molecules causes the variation of surface electrostatic potential of amines, and subsequently tailor the adsorption behavior of CO2. This is also confirmed by experimental results that the best performance is obtained on the adsorbent with comparable density of A and P. Because the chemical active sites are specific for amines, high selectivity of CO2 over CH4 is obtained, and such selectivity is tunable upon irradiation with UV/VIS light. The present strategy may open up a new way for the construction of smart adsorbents by taking advantage of the interplay between active sites and stimuli motifs, which is impossible or hard to realize by conventional approaches.

Page 8 of 10

ASSOCIATED CONTENT Supporting Information The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/×××××. 1 H NMR spectrum, IR spectra, TGA curves, SEM images, UV-vis spectra, gas adsorption isotherms, and further information (PDF) AUTHOR INFORMATION Corresponding Author *[email protected] Author Contributions ‡ Lei Cheng and Yao Jiang contributed equally. Notes The authors declare no competing financial interests. ACKNOWLEDGMENTS This work was supported by the National Natural Science Foundation of China (21722606, 21676138, and 21576137), the Project of Priority Academic Program Development of Jiangsu Higher Education Institutions, and the Postgraduate Research & Practice Innovation Program of Jiangsu Province (KYCX17_0964). REFERENCES (1) McFarland, E. Unconventional Chemistry for Unconventional Natural Gas. Science 2012, 338, 340-342. (2) Kotchen, M. J.; Mansur, E. T. Reassessing the Contribution of Natural Gas to US CO2 Emission Reductions Since 2007. Nat. Commun. 2016, 7, 10648. (3) Li, J. R.; Kuppler, R. J.; Zhou, H. C. Selective Gas Adsorption and Separation in Metal-Organic Frameworks. Chem. Soc. Rev. 2009, 38, 1477-1504. (4) Souther, S. Drawbacks to Natural Gas. Science 2013, 340, 141. (5) Yoon, J. W.; Chang, H.; Lee, S.-J.; Hwang, Y. K.; Hong, D.-Y.; Lee, S.-K.; Lee, J. S.; Jang, S.; Yoon, T.-U.; Kwac, K.; Jung, Y.; Pillai, R. S.; Faucher, F.; Vimont, A.; Daturi, M.; Ferey, G.; Serre, C.; Maurin, G.; Bae, Y.-S.; Chang, J.-S. Selective Nitrogen Capture by Porous Hybrid Materials Containing Accessible Transition Metal Ion Sites. Nat. Mater. 2017, 16, 526-531. (6) Huang, Z.; Sednek, C.; Urynowicz, M. A.; Jin, Y.; Igwe, U.; Li, S.; Huang, Z.; Urynowicz, M. A.; Jin, S.; Guo, H.; Wang, Q.; Fallgren, P.; Jin, S. Low Carbon Renewable Natural Gas Production from Coalbeds and Implications for Carbon Capture and Storage. Nat. Commun. 2017, 8, 568. (7) Choi, S.; Drese, J. H.; Jones, C. W. Adsorbent Materials for Carbon Dioxide Capture from Large Anthropogenic Point Sources. ChemSusChem 2009, 2, 796-854. (8) Su, F.; Lu, C. CO2 Capture from Gas Stream by Zeolite 13X Using a Dual-Column Temperature/Vacuum Swing Adsorption. Energy Environ. Sci. 2012, 5, 9021-9027. (9) Agarwal, A.; Biegler, L. T.; Zitney, S. E. A Superstructure-Based Optimal Synthesis of PSA Cycles For Post-Combustion CO2 Capture. AIChE J. 2010, 56, 1813-1828. (10)Aznar, E.; Oroval, M.; Pascual, L.; Murguia, J. R.; Martinez-Manez, R.; Sancenon, F. Gated Materials for On-Command Release of Guest Molecules. Chem. Rev. 2016, 116, 561-718. (11)Qu, D. H.; Wang, Q. C.; Zhang, Q. W.; Ma, X.; Tian, H. Photoresponsive Host-Guest Functional Systems. Chem. Rev. 2015, 115, 7543-7588. (12)Wang, Y.; Shim, M. S.; Levinson, N. S.; Sung, H. W.; Xia, Y. StimuliResponsive Materials for Controlled Release of Theranostic Agents. Adv. Funct. Mater. 2014, 24, 4206-4220.

8 ACS Paragon Plus Environment

Page 9 of 10 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

(13)Cobo, I.; Li, M.; Sumerlin, B. S.; Perrier, S. Smart Hybrid Materials by Conjugation of Responsive Polymers to Biomacromolecules. Nat. Mater. 2015, 14, 143-159. (14)Yanai, N.; Uemura, T.; Inoue, M.; Matsuda, R.; Fukushima, T.; Tsujimoto, M.; Isoda, S.; Kitagawa, S. Guest-to-Host Transmission of Structural Changes for Stimuli-Responsive Adsorption Property. J. Am. Chem. Soc. 2012, 134, 4501-4504. (15)Motornov, M.; Sheparovych, R.; Lupitskyy, R.; MacWilliams, E.; Hoy, O.; Luzinov, I.; Minko, S. Stimuli-Responsive Colloidal Systems from Mixed Brush-Coated Nanoparticles. Adv. Funct. Mater. 2007, 17, 2307-2314. (16)Xia, L. W.; Xie, R.; Ju, X. J.; Wang, W.; Chen, Q.; Chu, L. Y. NanoStructured Smart Hydrogels with Rapid Response and High Elasticity. Nat. Commun. 2013, 4, 2226. (17)Coskun, A. Great Expectations: Can Artificial Molecular Machines Deliver on Their Promise? Chem. Soc. Rev. 2012, 41, 19-30. (18)Stuart, M. A.; Huck, W. T.; Genzer, J.; Muller, M.; Ober, C.; Stamm, M.; Sukhorukov, G. B.; Szleifer, I.; Tsukruk, V. V.; Urban, M.; Winnik, F.; Zauscher, S.; Luzinov, I.; Minko, S. Emerging Applications of Stimuli-Responsive Polymer Materials. Nat. Mater. 2010, 9, 101-113. (19)Kume, S.; Nomoto, K.; Kusamoto, T.; Nishihara, H. Intramolecular Electron Arrangement with a Rotative Trigger. J. Am. Chem. Soc. 2009, 131, 14198-14199. (20)Mura, S.; Nicolas, J.; Couvreur, P. Stimuli-Responsive Nanocarriers for Drug Delivery. Nat. Mater. 2013, 12, 991-1003. (21)Ji, W.; Li, N.; Chen, D.; Qi, X.; Sha, W.; Jiao, Y.; Xu, Q.; Lu, J. Coumarin-Containing Photo-Responsive Nanocomposites for NIR Light-Triggered Controlled Drug Release via a Two-Photon Process. J. Mater. Chem. B 2013, 1, 5942-5949. (22)Huang, N.; Ding, X.; Kim, J.; Ihee, H.; Jiang, D. A Photoresponsive Smart Covalent Organic Framework. Angew. Chem. Int. Ed. 2015, 54, 8704-8707. (23)Hoshino, Y.; Imamura, K.; Yue, M.; Inoue, G.; Miura, Y. Reversible Absorption of CO2 Triggered by Phase Transition of Amine-Containing Micro- and Nanogel Particles. J. Am. Chem. Soc. 2012, 134, 1817718180. (24)Takahiro, M.; Kazushi, K.; Aida, T. A Self-Locking Molecule Operative with a Photoresponsive Key. J. Am. Chem. Soc. 2006, 128, 11600-11605. (25)Kundu, P. K.; Samanta, D.; Leizrowice, R.; Margulis, B.; Zhao, H.; Borner, M.; Udayabhaskararao, T.; Manna, D.; Klajn, R. LightControlled Self-Assembly of Non-Photoresponsive Nanoparticles. Nat. Chem. 2015, 7, 646-652. (26)Li, H.; Hill, M. R.; Doblin, C.; Lim, S.; Hill, A. J.; Falcaro, P. Visible Light Triggered CO2 Liberation from Silver Nanocrystals Incorporated Metal-Organic Frameworks. Adv. Funct. Mater. 2016, 26, 4815-4821. (27)Qin, L.; Gu, W.; Wei, J.; Yu, Y. Piecewise Phototuning of SelfOrganized Helical Superstructures. Adv. Mater. 2018, 30, 1704941. (28)Borah, P.; Sreejith, S.; Anees, P.; Menon, N. V.; Kang, Y.; Ajayaghosh, A.; Zhao, Y. Near-IR Squaraine Dye-Loaded Gated Periodic Mesoporous Organosilica for Photo-Oxidation of Phenol in a Continuous-Flow Device. Sci. Adv. 2015, 1, e1500390. (29)Chen, S.; Chen, L. J.; Yang, H. B.; Tian, H.; Zhu, W. Light-Triggered Reversible Supramolecular Transformations of Multi-Bisthienylethene Hexagons. J. Am. Chem. Soc. 2012, 134, 13596-13599. (30)Liu, Z.; Wang, H. I.; Narita, A.; Chen, Q.; Mics, Z.; Turchinovich, D.; Klaui, M.; Bonn, M.; Mullen, K. Photoswitchable MicroSupercapacitor Based on a Diarylethene-Graphene Composite Film. J. Am. Chem. Soc. 2017, 139, 9443-9446. (31)Jochum, F. D.; Theato, P. Temperature- and Light-Responsive Smart Polymer Materials. Chem. Soc. Rev. 2013, 42, 7468-7483. (32)Fan, C. B.; Le Gong, L.; Huang, L.; Luo, F.; Krishna, R.; Yi, X. F.; Zheng, A. M.; Zhang, L.; Pu, S. Z.; Feng, X. F.; Luo, M. B.; Guo, G. C. Significant Enhancement of C2H2/C2H4 Separation by a Photochromic Diarylethene Unit: A Temperature- and Light-Responsive Separation Switch. Angew. Chem. Int. Ed. 2017, 56, 7900-7906. (33)Furlong, B. J.; Katz, M. J. Bistable Dithienylethene-Based MetalOrganic Framework Illustrating Optically Induced Changes in Chemical Separations. J. Am. Chem. Soc. 2017, 139, 13280-13283. (34)Huang, R.; Hill, M. R.; Babarao, R.; Medhekar, N. V. CO2 Adsorption in Azobenzene Functionalized Stimuli Responsive Metal-Organic Frameworks. J. Phys. Chem. C 2016, 120, 16658-16667.

(35)Knebel, A.; Sundermann, L.; Mohmeyer, A.; Strauß, I.; Friebe, S.; Behrens, P.; Caro, J. Azobenzene Guest Molecules as Light-Switchable CO2 Valves in an Ultrathin UiO-67 Membrane. Chem. Mater. 2017, 29, 3111-3117. (36)Huang, H.; Sato, H.; Aida, T. Crystalline Nanochannels with Pendant Azobenzene Groups: Steric or Polar Effects on Gas Adsorption and Diffusion? J. Am. Chem. Soc. 2017, 139, 8784-8787. (37)Li, H.; Martinez, M. R.; Perry, Z.; Zhou, H.-C.; Falcaro, P.; Doblin, C.; Lim, S.; Hill, A. J.; Halstead, B.; Hill, M. R. A Robust Metal-Organic Framework for Dynamic Light-Induced Swing Adsorption of Carbon Dioxide. Chem. - Eur. J. 2016, 22, 11176-11179. (38)Zhu, Y.; Zhang, W. Reversible Tuning of Pore Size and CO2 Adsorption in Azobenzene Functionalized Porous Organic Polymers. Chem. Sci. 2014, 5, 4957-4961. (39)Li, H.; Sadiq, M. M.; Suzuki, K.; Doblin, C.; Lim, S.; Falcaro, P.; Hill, A. J.; Hill, M. R. MaLISA-a Cooperative Method to Release Adsorbed Gases from Metal-Organic Frameworks. J. Mater. Chem. A 2016, 4, 18757-18762. (40)L., G. L.; Feng, X. F.; Luo, F. Novel azo-Metal-Organic Framework Showing a 10-Connected bct Net, Breathing Behavior, and Unique Photoswitching Behavior toward CO2. Inorg. Chem. 2015, 54, 1158711589. (41)Park, J.; Yuan, D.; Pham, K. T.; Li, J. R.; Yakovenko, A.; Zhou, H. C.; Zhou, H. C. Reversible Alteration of CO2 Adsorption Upon Photochemical or Thermal Treatment in a Metal-Organic Framework. J. Am. Chem. Soc. 2012, 134, 99-102. (42)Lyndon, R.; Konstas, K.; Ladewig, B. P.; Southon, P. D.; Kepert, P. C.; Hill, M. R. Dynamic Photo-Switching in Metal-Organic Frameworks as a Route to Low-Energy Carbon Dioxide Capture and Release. Angew. Chem. Int. Ed. 2013, 52, 3695-3698. (43)Lee, J. W.; Klajn, R. Dual-Responsive Nanoparticles that Aggregate Under the Simultaneous Action of Light and CO2. Chem. Commun. 2015, 51, 2036-2039. (44)Manna, D.; Udayabhaskararao, T.; Zhao, H.; Klajn, R. Orthogonal Light-Induced Self-Assembly of Nanoparticles using Differently Substituted Azobenzenes. Angew. Chem., Int. Ed. 2015, 54, 1239412397. (45)Liu, N.; Chen, Z.; Dunphy, D. R.; Jiang, Y. B.; Assink, R. A.; Brinker, C. J. Photoresponsive Nanocomposite Formed by Self-Assembly of an Azobenzene-Modified Silane. Angew. Chem. 2003, 42, 1731-1734. (46)Tan, S. Y.; Ang, C. Y.; Li, P.; Yap, Q. M.; Zhao, Y. Drug Encapsulation and Release by Mesoporous Silica Nanoparticles: The Effect of Surface Functional Groups. Chemistry 2015, 20, 11276-11282. (47)Rowsell, J. L.; 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)Myers, A. L.; Prausnitz, J. M. Thermodynamics of Mixed-Gas Adsorption. AIChE J. 2010, 11, 121-127. (49)Chen, T.; Chen, R.; Zhong, J.; Liu, J. Engineering Hollow Mesoporous Silica Nanocontainers with Molecular Switches for Continuous SelfHealing Anticorrosion Coating. J. Mater. Chem. A 2015, 3, 9510-9516. (50)Adam, A.; Haberhauer, G. Switching Process Consisting of Three Isomeric States of an Azobenzene Unit. J. Am. Chem. Soc. 2017, 139, 9708-9713. (51)Tanaka, T.; Ogino, H.; Iwamoto, M. Photochange in Pore Diameters of Azobenzene-Planted Mesoporous Silica Materials. Langmuir 2007, 23, 11417-11420. (52)Wang, Z.; Grosjean, S.; Brase, S.; Heinke, L. Photoswitchable Adsorption in Metal-Organic Frameworks Based on Polar Guest-Host Interactions. ChemPhysChem 2015, 16, 3779-3783. (53)Zhao, H.; Sen, S.; Udayabhaskararao, T.; Sawczyk, M.; Kucanda, K.; Manna, D.; Kundu, P. K.; Lee, J. W.; Kral, P.; Klajn, R. Reversible Trapping and Reaction Acceleration within Dynamically SelfAssembling Nanoflasks. Nat. Nanotechnol. 2016, 11, 82-88.

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

Page 10 of 10

For Table of Contents Only

10 ACS Paragon Plus Environment