Surface Pore Engineering of Covalent Organic Frameworks for

Jun 6, 2018 - Figure 1. (a) Scheme for surface pore engineering of COFs with various groups and (b) ... In this context, we use TpPa-1 ([HOOC]0-COF) m...
0 downloads 0 Views 3MB Size
This is an open access article published under an ACS AuthorChoice License, which permits copying and redistribution of the article or any adaptations for non-commercial purposes.

Research Article Cite This: ACS Cent. Sci. XXXX, XXX, XXX−XXX

Surface Pore Engineering of Covalent Organic Frameworks for Ammonia Capture through Synergistic Multivariate and Open Metal Site Approaches Yajie Yang,† Muhammad Faheem,† Lili Wang,† Qinghao Meng,† Haoyan Sha,‡ Nan Yang,§ Ye Yuan,*,† and Guangshan Zhu*,† †

Key Laboratory of Polyoxometalate Science of Ministry of Education, Northeast Normal University, Changchun 130024, P. R. China Department of Chemical Engineering, University of California, Davis, Davis, California 95616, United States § China Faw New Energy Vehicle Branch, Changchun 130011, P. R. China ‡

S Supporting Information *

ABSTRACT: Ammonia (NH3) is a commonly used industrial gas, but its corrosiveness and toxicity are hazardous to human health. Although many adsorbents have been investigated for NH3 sorption, limited ammonia uptake remains an urgent issue yet to be solved. In this article, a series of multivariate covalent organic frameworks (COFs) are explored which are densely functionalized with various active groups, such as  NH, CO, and carboxyl group. Then, a metal ion (Ca2+, Mn2+, and Sr2+) is integrated into the carboxylated structure achieving the first case of an open metal site in COF architecture. X-ray photoelectron spectroscopy reveals conclusive evidence for the multiple binding interactions with ammonia in the modified COF materials. Infrared spectroscopy indicates a general trend of binding capability from weak to strong along with NH, CO, carboxyl group, and metal ion. Through the synergistic multivariate and open metal site, the COF materials show excellent adsorption capacities (14.3 and 19.8 mmol g−1 at 298 and 283 K, respectively) and isosteric heat (Qst) of 91.2 kJ mol−1 for ammonia molecules. This novel approach enables the development of tailor-made porous materials with tunable pore-engineered surface for ammonia uptake. fields of catalysis, sensing, optoelectricity, and gas adsorption and separation.16−20 Therefore, one can create tailor-made binding sites through suitable selection of building block or modification strategies for target functions. Desirable properties, including extensive pore surface area and tunable surface chemistry, attract great interest for the development of novel COF-based adsorbents for NH3 sorption. Herein, we present a series of multivariate COF materials to integrate various binding groups into the porous network for surface pore engineering. A three-component system comprising triformylphloroglucinol (TFP), 2,5-diaminobenzoic acid (DAA), and p-phenylenediamine (PA-1) at different molar ratios (X = [DAA]/([DAA] + [PA − 1]) × 100 = 0, 17, 33, 50, and 100) is explored to prepare [HOOC]X-COFs (X = 0, 17, 33, 50, and 100), as shown in Figure 1. The [HOOC]17-COF, with its remarkable NH3 adsorption capacity, was then poured into the chloride salt to anchor Lewis centers (Ca2+, Mn2+, and Sr2+) onto the pore surface. The metallized COF product first obtains the open metal site to enhance the chemical adsorption affinity for NH3 molecules by forming acidic−basic adducts. Xray photoelectron spectroscopy (XPS) and infrared (IR)

1. INTRODUCTION Industrial development has led to a great demand for ammonia as a chief raw material, with annual production exceeding 200 million tons.1 However, its strong causticity and toxicity give rise to the identification of threats from industrial emissions and spills, which can cause convulsions, coma, and death to all vertebrates.2 To eliminate such hazards, engineered materials that can adsorb large amounts of ammonia are desperately needed. Although numerous adsorbents, such as activated carbons, zeolites, and metal−organic frameworks (MOFs), have shown potential for applications in NH3 sorption, the progressive performance of ammonia uptake remains a challenging issue.3−7 To address this objective, two qualities are considered as vital ingredients for an efficient adsorbent:8 (1) a highly porous framework consisting of unique microsized pores that can increase the physical adsorption of ammonia; and (2) effective functional groups in the porous structure that can interact with ammonia and favor chemical adsorption. Covalent organic frameworks (COFs) are an emerging class of crystalline porous materials that are composed of light elements (e.g., H, B, C, N, O) via covalent bonds.9−15 Using state-of-the-art molecular design, COF-based structures can exhibit a large surface area, unique porosity, high crystallinity, and tunable pore chemistry, with promising applications in the © XXXX American Chemical Society

Received: April 18, 2018

A

DOI: 10.1021/acscentsci.8b00232 ACS Cent. Sci. XXXX, XXX, XXX−XXX

Research Article

ACS Central Science

Figure 1. (a) Scheme for surface pore engineering of COFs with various groups and (b) possible pore structure of COFs with various groups (gray, C; blue, N; red, O; yellow, metal).

spectroscopy. As shown in Figure S2, the consumption of starting materials can be observed from the appearance of  NH bending at 1520 cm−1 and CO stretching at 1625 cm−1 in the COF skeleton. The main bands at 989 and 1657 cm−1 are assigned to the plane bending vibration of OH and stretching vibration of CO in the carboxyl group, respectively. CHN elemental analysis reveals the contents of the carboxyl group in [HOOC]X-COFs (X = 0, 17, 33, 50, and 100), with the measured result in concert with theoretical calculations (Table S1). The architectural stability is manifested by thermogravimetric (TG) analysis, with no weight loss found for each COF until decomposition at 250 °C (Figure S3). Powder X-ray diffraction (PXRD) patterns show the same ordered structure as that found for TpPa-1, but the crystallinity of [HOOC]X-COF (X = 0, 17, 33, 50, and 100) decreases for increasing X, as shown in Figure S4. Scanning electron microscopy (SEM) images show that [HOOC]X-COFs (X = 0, 17, 33, 50, and 100) are crystallized as microsized (1−5 μm) aggregated particles (Figure S5). The transmission electron microscopy (TEM) images show that some of the COF samples exhibit a sheetlike structure due to π−π stacking of the COF layers (Figure S6). The porosity of the COFs was measured by N2 adsorption isotherms for the fully activated samples. The surface areas for the [HOOC]X-COFs with carboxyl group content (X) varying from 0 to 17, 33, 50, and 100 was calculated by the BET model to be 713, 652, 458, 279, and 150 m2 g−1, respectively (Figure S7). This descending trend in the surface area is due to the electron withdrawing efficiency of the carboxyl group, which reduces the activity of the coupling reaction during the polymerization process leading to ruptures in the integration of the COF structure.27,28 The pore sizes varied from 1.50 to 1.48, 1.44, 1.11, and 0.97 nm as the amount of carboxyl group

spectroscopy provided solid evidence of the multiple interactions with ammonia, including the formation of hydrogen bonds with NH and CO and acidic− basic adducts with carboxyl group and metal ions. Due to the synergistic effect of multivariate components and open metal site, the COF materials exhibited considerable total adsorption capacities for ammonia realized via multiple binding interactions.

2. RESULTS AND DISCUSSION In this context, we use TpPa-1 ([HOOC]0-COF) material as a scaffold21,22 which features a well-defined hexagonal structure with space group (P6/m), high BET surface area of 713 m2 g−1, and unique pore size distribution (PSD) centered at 1.50 nm calculated by the nonlocal−density functional theory (NL− DFT) method. Two particularly salient features of [HOOC]0COF suggest that it will be an excellent scaffold for the development of novel ammonia adsorbents. (1) The unique microchannel in the porous skeleton favors physical adsorption of ammonia via van der Waals force.21 (2) Diverse fragments, including NH and CO group, can serve as protondonating or proton-accepting groups for hydrogen-bonding interactions, which can significantly enhance the affinity to ammonia.23−26 As illustrated in Figure 2 and Figure S1, [HOOC]0-COF shows a meaningful NH3 (2.56 Å) capacity of 6.85 and 9.23 mmol g−1 at 298 and 283 K, respectively. The isosteric heat of NH3 adsorption based on the Clausius− Clapeyron relation is determined to be 43.5 kJ mol−1 from independent fits to the 283 and 298 K isotherms. Compared with [HOOC]0-COF, the carboxyl group is introduced into the porous skeleton to produce the multivariate COFs by use of different ratios of linkers. The presence of the carboxyl group in [HOOC]X-COFs was confirmed by FT-IR B

DOI: 10.1021/acscentsci.8b00232 ACS Cent. Sci. XXXX, XXX, XXX−XXX

Research Article

ACS Central Science

a larger storage capacity by chemical adsorption. After immersing into each chloride salt solution, a series of [MOOC]17-COF materials were obtained: [CaOOC]17-COF, [MnOOC]17-COF, and [SrOOC]17-COF. Infrared spectroscopy (Figure S2) for these materials presents weak bands in the scope of 502−585 and 444−472 cm−1, which are ascribed to the stretching vibrations of MO and MN, respectively.30 The existence of OH signal at 989 cm−1 and a reduced intensity for the CO band in the carboxyl group (1657 cm−1) of [MOOC]17-COFs suggest the coordination of C O and NH groups to the metal ion (Figure 1). Based on energy minimization optimization calculated by the Materials Studio (MS) simulation, two O (from CO of COOH and COF skeleton) and one N (from NH) atoms coordinate to one metal ion in the [MOOC]17-COF framework to form a stabilized double six-membered ring structure (Figure 1), which is consistent with the conclusion of a previous investigation.31 Elemental analysis reveals the almost equal content of metal ion and COOH group in [MOOC]17-COFs (Ca2+, Mn2+, and Sr2+), which is consistent with the result of the theoretical model (Table S1). The weight losses of 2.7%, 7.3%, and 4.5% before 100 °C are ascribed to the loss of guest molecules (Figure S3). After heating samples up to 800 °C, ca. 3.2%, 4.8%, and 5.9% of residues remain corresponding to the [MOOC]17-COFs (Ca2+, Mn2+, and Sr2+), which is in agreement with the results of the elemental analysis. PXRD patterns reveal that the [MOOC]17-COFs (Ca2+, Mn2+, and Sr2+) retain the same crystallinity as the [HOOC]17-COF even after metal ion incorporation (Figure S4). As shown in Figure S5, SEM images for the [MOOC]17-COF (Ca2+, Mn2+, and Sr2+) also show aggregated particles. The metal ions are uniformly dispersed in the COF structure, as indicated by no metallic nanoparticles visible in the TEM images (Figure S6). The BET surface areas are determined to be 629, 614, and 587 m 2 g −1 for [CaOOC] 17 -COF, [MnOOC] 17 -COF, and [SrOOC]17-COF, respectively (Figure S9). This explicit decrease in area results from the molecular weight being scaled up along with incorporation of the Ca2+, Mn2+, and Sr2+ ions. Upon introduction of the Ca2+, Mn2+, or Sr2+ ions onto the pore surface, the resulting [MOOC]17-COFs with open metal sites present higher sorption amounts (at 298 K) of 12.25, 11.38, and 14.30 mmol g−1, respectively. The strong affinity of open metal site is also demonstrated by the sharp increase of uptake capacity at low pressure range (Figure S10). Likewise, after the desorption processes, the three COF materials exhibit a hysteresis terminating with uptakes of 1.53, 0.85, and 1.99 mmol g−1 for [CaOOC]0-COF, [MnOOC]17-COF, and [SrOOC]17-COF, respectively. After calculation, we speculate that each Ca2+, Mn2+, and Sr2+ ion can uphold 2.01, 1.69, and 2.65 NH3 molecules, respectively. The Sr2+ ion possesses the highest uptake capability because of its specific electronic structure, which favors formation of a multiple (a maximum of 6) coordinated complex with NH3 molecules.32−35 After interaction with NH3 molecules, the tricoordinated Sr2+ ion in the double six-membered ring structure can combine with three more ammonia molecules to reach its maximum coordination structure, which is basically identical to the experimental result of 2.65 NH3 molecules. [SrOOC]17-COF possesses an ultrahigh capacity of 19.8 mmol g−1 at 283 K and 1 bar (Figure S11). After introduction of the Sr2+ ion, the [SrOOC]17-COF shows a significant improvement in the isosteric heat for NH3 gas (91.2 kJ mol−1).

Figure 2. NH3 adsorption (solid symbols) and desorption (open symbols) at 1 bar and 298 K for activated samples of [HOOC]0-COF (a), [HOOC]17-COF (b), [HOOC]33-COF (c), [HOOC]50-COF (d), [HOOC]100-COF (e), [CaOOC]17-COF (f), [MnOOC]17-COF (g), and [SrOOC]17-COF (h).

increased from 0 to 17, 33, 50, and 100, respectively (Figure S7). [HOOC]X-COFs with X values of 17, 33, 50, and 100 showed an NH3 capacity of 9.34, 8.21, 6.67, and 4.14 mmol g−1, respectively, at 298 K and 1 bar (Figure 2). These decreased NH3 uptakes are attributed to the reduced porous surfaces and lack of long-range order in the [HOOC]X-COFs. The [HOOC]17-COF was selected for further investigations despite having a lower surface area (652 m2 g−1) compared to [HOOC]0-COF (713 m2 g−1); we believe this COF should show a strong attractive interaction between the carboxyl group (−COOH) and NH3 molecule,8,29 this hypothesis is confirmed by the enhanced isosteric heat of NH3 adsorption (55.3 kJ mol−1) measured at 298 and 283 K (Figure S8). Therefore, this result clearly demonstrates the vital role of the additional functional group (−COOH) in the COF pore surface. The [HOOC]17-COF with the best NH3 uptake is selected as a subject for further surface engineering to improve the performance. Previous studies reported that Lewis centers, such as Ca2+, Mn2+, and Sr2+ ions on the surface of a porous framework, may create a strong affinity for NH3 molecules.3−8 Therefore, respective metal ions are incorporated into the [HOOC]17-COF architecture to achieve the open metal site for C

DOI: 10.1021/acscentsci.8b00232 ACS Cent. Sci. XXXX, XXX, XXX−XXX

Research Article

ACS Central Science

Figure 3. XPS spectra for [HOOC]0-COF (a, b), [HOOC]17-COF (c, d), and [SrOOC]17-COF (e−g). Black curves represent activated samples, and red curves represent exhausted samples following interaction with NH3.

[SrOOC]17-COF reveals a decent capacity (10.92 mmol g−1) after three consecutive adsorption/desorption cycles (Figure S12). Such a decrease in NH3 uptake is due to strong acid−base interactions that unite for a certain amount of ammonia in the framework, limiting desorption under vacuum conditions. The TGA result indicates that [SrOOC]17-COF will desorb ammonia molecules at ∼150 °C (Figure S13) while retaining its crystalline structure for cycle use. The bound ammonia calculated from ca. 3.4% weight loss among 120−170 °C in the TGA curve also agrees with the sorption calculation that each Sr2+ ion can uphold 2.73 NH3 molecules. After heated at 200 °C under vacuum for 12 h, [SrOOC]17-COF would release the bound ammonia molecules and reabsorb ca. 14.12 mmol g−1 NH3 molecules (Figure S14). The capacity decreased by only 7% after five cycles indicating that [SrOOC]17-COF exhibits good reusability (Figure S15). The large adsorption capacity of [SrOOC]17-COF (∼14.30 mmol g−1 at 298 K) is much higher than those of most other adsorbents such as MCM-41 (7.9 mmol g−1), 13X zeolite (9.0 mmol g−1), Amberlyst 15 (11.0 mmol g −1 ), Co 2 Cl 2 (BTDD)(H 2 O) 2 (12.0 mmol g −1 ), Ni2Cl2(BTDD)(H2O)2 (12.0 mmol g−1), and PPN-6-SO3H (12.1 mmol g−1) and is close to the highest level of ammonia adsorbents, including Mn2Cl2(BTDD)-(H2O)2 (15.5 mmol g−1), BPP-5 (17.7 mmol g−1), and Cu2Cl2BBTA (19.79 mmol g−1).3−8,10 Usually, the absorptivity of a porous material is related to the surface area and pore volume; however, none of these correlations hold in our comparative investigations. For comparison, the classical porous aromatic framework (PAF-1) with ultrahigh surface area (4240 m2 g−1 based on a BET

model) and uniform pore channel (1.41 nm) was demonstrated to contrast the impacts of porosity and aromatic rings.36 With similar pore size but larger surface area, PAF-1 exhibits ∼0.2 mmol g−1 capacity at 298 K.8 It is evident that the adsorption capability of COF materials in this work is highly dependent on the synergistic effect of multivariate components and open metal site. XPS study affords a surface analytical technique that can be used to accurately determine the chemical state of each element. [HOOC]0-COF, [HOOC]17-COF, and [SrOOC]17COF samples for XPS were prepared by first heating to 473 K for 24 h under vacuum, followed by exposure of the activated COFs to a dry NH3 environment for 72 h at 283 K (Figure 3). As illustrated in Figure 3, ammonia with a nitrogen 1s binding energy of 397.9 eV could not be isolated possibly because of the weak signal of gas molecules.37 The binding energies for N 1s and O 1s in the [HOOC]0-COF center are at 400.0 and 532.3 eV, respectively (Figure 3a,b). After NH3 (∼397.9 eV) gas adsorption, these binding energies change to ca. 399.6 eV for N 1s and 532.0 eV for O 1s, which manifests the barely exposed H nuclei in NH3 via the hydrogen bond that pulls the electron cloud of the donor atom (N and O) of the framework resulting in a decrease of the binding energies.38,39 In [HOOC]17-COF (Figure 3c), N 1s photoemission in the activated sample occurs at 400.0 eV and is decreased to 399.6 eV after exposure to an NH3 environment due to the formation of hydrogen bonds. A shoulder peak for N 1s is located at 401.0 eV due to protonation of NH3, which can be used to identify the formation of the ammonium (NH4+) salt.40 The reducing binding energies at 513.5 and 533.2 eV can be ascribed to the O atoms in CO and COOH, respectively (Figure 3d). D

DOI: 10.1021/acscentsci.8b00232 ACS Cent. Sci. XXXX, XXX, XXX−XXX

Research Article

ACS Central Science

Figure 4. NH3 adsorption against different temperatures at 1 bar for activated samples of [HOOC]0-COF (a), [HOOC]17-COF (b), and [SrOOC]17-COF (c).

temperature for the exhausted sample was monitored by an infrared detector until the temperature reached 423 K. From the IR spectroscopy, several features are observed as the exhausted sample desorbs NH3 from the framework at different temperatures. Here, the CN bond (1255 cm−1) is selected as the internal standard; variation in peak intensity is confirmed based on the ratio of the intensity of the characteristic bond to the CN. As observed in Figure 4a, the value (relative intensity for the NH bond compared to the CN bond) increases from 1.0 in the original [HOOC]0-COF to ca. 1.2 for the adsorbed-ammonia COF material (283 K), which indicates interaction between the  NH group and ammonia molecule.45 Moreover, the ratio changes back to a value of 1.0 after the sample is heated up to 303 K. The data show a wide band (1630 cm−1) for CO in the [HOOC]0 -COF structure; following adsorption of ammonia at 283 K, the location of the maximum for the  CO band toward lower wavenumber (1614 cm−1) can be attributed to the formation of hydrogen bonds. With increasing temperature to 323 K, desorption of ammonia leads to reversion of the band position back to 1630 cm−1. Similarly, we utilize the FT-IR spectra to probe the interactions between [HOOC]17-COF and the ammonia adsorbate before and after adsorption (Figure 4b). In addition to the two characteristics for [HOOC]0-COF, a new feature at 1430 cm−1 is observed for [HOOC]17-COF exposed to ammonia, which can be assigned to the vibration of NH in NH4+.46,47 This observation suggests that a fraction of the ammonia molecules are converted into the ionic form via the acid−base reaction with the carboxylic group. This band disappeared at 363 K, indicating the removal of ammonia. As to the [SrOOC]17COF (Figure 4c), another band at 1496 cm−1 is ascribed to the existence of a scissoring mode for the amide species, NH2,

The high-resolution O 1s spectrum observed at 532.3 eV is resolved into two characteristic peaks for CO in H bond and COO− in the ammonium salt. For [SrOOC]17-COF (Figure 3e), the N 1s peak component at 400.0 eV is shifted by −0.4 eV compared to the reacted N atom, which is clearly evident of hydrogen bond formation, with protonation of NH3 also shown at 400.8 eV because of the formation of the ammonium salt. Coordination of the O atom to the Sr2+ ion is demonstrated by the lower shifted peak (530.9 eV) for C O and the weakened intensity (533.2 eV) for the COOH (Figure 3f). The N 1s and O 1s changes for [SrOOC]17-COF in XPS spectra agree with the formation of the stabilized double six-membered ring structure. As for the exhausted [SrOOC]17COF, the strong signal observed at 532.3 eV fits perfectly with the two characteristic peaks corresponding to CO in H bond and COO− in the ammonium salt. It is evident from the data shown in Figure 3g that a coordinated structure for Sr2+ with ligand atoms is formed, as the binding energy of Sr 3d reducing from 134.5 eV (SrCl2)41 to 134.2 eV ([SrOOC]17COF). This energy shift is due to capable elements clinging to the Sr2+ surface, and withdrawal of the electron cloud from the Sr atom to the coordinated O and N atoms themselves, leading to an enhanced shielding effect that decreases the binding energy.42−44 After the coordination of NH3 to Sr2+ ion, the electron cloud of the Sr atom is pulled further toward the N atoms, resulting in a decrease in the binding energy for the Sr atom (133.7 eV) in the NH3-adsorbed [SrOOC]17-COF. The effect of each group is assessed by IR spectroscopy at different temperatures to qualitatively indicate the affinity (Figure 4). After adsorption of NH3 molecules, the respective COF samples were heated from 283 K, with the temperature increased in 10 degree steps and constant holding of the temperature for 10 min before further heating. The effect of E

DOI: 10.1021/acscentsci.8b00232 ACS Cent. Sci. XXXX, XXX, XXX−XXX

ACS Central Science



which is formed from the coordination of the Sr2+ ion and ammonia.48,49 The relative intensity of the NH2 bond decreases gradually from 393 K and ultimately disappears at 423 K. One can speculate that the binding capability of N H, CO, carboxyl group, and metal ion trends from weak to strong.

Research Article

ASSOCIATED CONTENT

* Supporting Information S

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acscentsci.8b00232. Linker syntheses and characterization, PXRD patterns, TGA curves, NMR spectroscopy, and adsorption isotherms (PDF)



3. CONCLUSIONS In summary, surface pore engineering of COF materials is explored via the incorporation of various functional units on pore walls. Due to the synergistic multivariate and open metal site, the porous architecture profoundly enhances the binding affinity for ammonia molecules. Notably, the role and tendency of the NH, CO, COOH, and metal ions in interactions with ammonia can guide the functionalization of other porous materials for separation and adsorption of ammonia. Additionally, we expect our approach based on the “surface pore engineering” will motivate research on utilizing COF materials in other fields, with further progress eventually leading to industrial applications.

AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected]. *E-mail: [email protected]. ORCID

Guangshan Zhu: 0000-0002-5794-3822 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS We are grateful for financial support from the National Basic Research Program of China (973 Program, Grants 2012CB821700, 2014CB931800), the National Natural Science Foundation of China (NSFC Grants 21401069, 21501024, 91622106, 21604008), the Major International (Regional) Joint Research Project of NSFC (Grant 21120102034), and Natural Science Foundation of Jilin Province of China (Grant 20180520144JH).

4. EXPERIMENTAL SECTION Synthesis of [HOOC]X-COFs. 2,5-Diaminobenzoic acid (DAA) and p-phenylenediamine (PA-1) were added into a solution of triformylphloroglucinol (TFP) (63 mg, 0.3 mmol), mesitylene (1.5 mL), dioxane (1.5 mL), and 3 M aqueous acetic acid (0.5 mL). Molar ratios for DAA and PA-1 of 0:6, 1:5, 2:4, 3:3, and 6:0 in total amount of 0.9 mmol were used to produce [HOOC]0-COF, [HOOC]17-COF, [HOOC]33-COF, [HOOC]50-COF, and [HOOC]100-COF, respectively. The mixture was then frozen by using a liquid N2 bath (77 K) and degassed via three freeze−pump−thaw cycles. After heating at 120 °C for 3 days, the precipitate was collected by filtration and washed with copious amounts of anhydrous dimethylacetamide (DMAC) and acetone. Finally, the resulting red powder was dried under vacuum at 180 °C for 24 h yielding a series of [HOOC]X-COF samples. Preparation of [MOOC]17-COFs. A 0.100 g portion of activated [HOOC]17-COF with the −COOH group (0.65 mmol g−1) was poured into a chloride salt (CaCl2, MnCl2, and SrCl2) aqueous solution (15.6 mmol L−1, 5 mL), respectively. After the solution was stirred for 24 h, the insoluble product was filtered out. Then, the precipitate was thoroughly washed with water three times to afford [MOOC]17-COFs (Ca2+, Mn2+, and Sr2+). NH 3 Release Experiment against Temperature. [HOOC]0-COF, [HOOC]17-COF, and [SrOOC]17-COF were heated under vacuum to approximately 473 K for 12 h and then cooled to room temperature. Next, the activated COFs were exposed to a dry NH3 environment for 24 h at 283 K. The respective COF samples were then heated to 283 K, with the temperature increased 10 degrees at a time and held constant for 10 min at each step before further heating. The effect of temperature for the exhausted sample was monitored by an infrared detector until the temperature reached 423 K. General experiments for the complete physical characterization of all as-synthesized COF materials are available in the Supporting Information. No unexpected or unusually high safety hazards were encountered.



REFERENCES

(1) DeCoste, J. B.; Denny, M. S.; Peterson, G. W.; Mahle, J. J.; Cohen, S. M. Enhanced aging properties of HKUST-1 in hydrophobic mixed-matrix membranes for ammonia adsorption. Chem. Sci. 2016, 7, 2711. (2) Randall, D. J.; Tsui, T. K. N. Ammonia toxicity in fish. Mar. Pollut. Bull. 2002, 45, 17. (3) Petit, C.; Bandosz, T. J. Enhanced adsorption of ammonia on metal-organic framework/graphite oxide composites: analysis of surface interactions. Adv. Funct. Mater. 2010, 20, 111. (4) Bobbitt, N. S.; Mendonca, M. L.; Howarth, A. J.; Islamoglu, T.; Hupp, J. T.; Farha, O. K.; Snurr, R. Q. Metal-organic frameworks for the removal of toxic industrial chemicals and chemical warfare agents. Chem. Soc. Rev. 2017, 46, 3357. (5) Petit, C.; Kante, K.; Bandosz, T. J. The role of sulfur-containing groups in ammonia retention on activated carbons. Carbon 2010, 48, 654. (6) Rieth, A. J.; Tulchinsky, Y.; Dincă, M. High and reversible ammonia uptake in mesoporous azolate metal−organic frameworks with open Mn, Co, and Ni Sites. J. Am. Chem. Soc. 2016, 138, 9401. (7) Rieth, A. J.; Dincă, M. Controlled gas uptake in metal−organic frameworks with record ammonia sorption. J. Am. Chem. Soc. 2018, 140, 3461. (8) Van Humbeck, J. F.; McDonald, T. M.; Jing, X.; Wiers, B. M.; Zhu, G.; Long, J. R. Ammonia capture in porous organic polymers densely functionalized with Bronsted acid groups. J. Am. Chem. Soc. 2014, 136, 2432. (9) Huang, N.; Wang, P.; Jiang, D. Covalent organic frameworks: a materials platform for structural and functional designs. Nat. Rev. Mater. 2016, 1, 16068. (10) Doonan, C. J.; Tranchemontagne, D. J.; Glover, T. G.; Hunt, J. H.; Yaghi, O. M. Exceptional ammonia uptake by a covalent organic framework. Nat. Chem. 2010, 2, 235. (11) Calik, M.; Auras, F.; Salonen, L. M.; Bader, K.; Grill, I.; Handloser, M.; Medina, D. D.; Dogru, M.; Löbermann, F.; Trauner, D.; Hartschuh, A.; Bein, T. Extraction of photogenerated electrons and holes from a covalent organic framework integrated heterojunction. J. Am. Chem. Soc. 2014, 136, 17802. F

DOI: 10.1021/acscentsci.8b00232 ACS Cent. Sci. XXXX, XXX, XXX−XXX

Research Article

ACS Central Science (12) Ding, S.-Y.; Dong, M.; Wang, Y.-W.; Chen, Y.-T.; Wang, H.-Z.; Su, C.-Y.; Wang, W. Thioether-Based Fluorescent Covalent Organic Framework for Selective Detection and Facile Removal of Mercury(II). J. Am. Chem. Soc. 2016, 138, 3031. (13) Medina, D. D.; Rotter, J. M.; Hu, Y.; Dogru, M.; Werner, V.; Auras, F.; Markiewicz, J. T.; Knochel, P.; Bein, T. Room temperature synthesis of covalent−organic framework films through vapor-assisted conversion. J. Am. Chem. Soc. 2015, 137, 1016. (14) Zhang, J.; Han, X.; Wu, X.; Liu, Y.; Cui, Y. Multivariate Chiral Covalent Organic Frameworks with Controlled Crystallinity and Stability for Asymmetric Catalysis. J. Am. Chem. Soc. 2017, 139, 8277. (15) Lin, S.; Diercks, C. S.; Zhang, Y.; Kornienko, N.; Nichols, E. M.; Zhao, Y.; Paris, A. R.; Kim, D.; Yang, P.; Yaghi, O. M.; Chang, C. J. Covalent organic frameworks comprising cobalt porphyrins for catalytic CO2 reduction in water. Science 2015, 349, 1208. (16) Wang, X.; Maeda, K.; Thomas, A.; Takanabe, K.; Xin, G.; Carlsson, J. M.; Domen, K.; Antonietti, M. A metal-free polymeric photocatalyst for hydrogen production from water under visible light. Nat. Mater. 2009, 8, 76. (17) Sun, Q.; Fu, C.-W.; Aguila, B.; Perman, J. A.; Wang, S.; Huang, H.-Y.; Xiao, F.-S.; Ma, S. Pore environment control and enhanced performance of enzymes infiltrated in covalent organic frameworks. J. Am. Chem. Soc. 2018, 140, 984. (18) Diercks, C. S.; Yaghi, O. M. Parallels between framework and molecular assembly. Science 2017, 355, 923. (19) Sun, Q.; Aguila, B.; Earl, L. D.; Abney, C. W.; Wojtas, L.; Thallapally, P. K.; Ma, S. Covalent Organic Frameworks as a Decorating Platform for Utilization and Affinity Enhancement of Chelating Sites for Radionuclide Sequestration. Adv. Mater. 2018, 30, 1705479. (20) Wu, D.; Xu, F.; Sun, B.; Fu, R.; He, H.; Matyjaszewski, K. Design and preparation of porous polymers. Chem. Rev. 2012, 112, 3959. (21) Kandambeth, S.; Mallick, A.; Lukose, B.; Heine, T.; Banerjee, R. Construction of crystalline 2D covalent organic frameworks with remarkable chemical (acid/base) stability via a combined reversible and irreversible route. J. Am. Chem. Soc. 2012, 134, 19524. (22) Biswal, B. P.; Kandambeth, S.; Chandra, S.; Shinde, D. B.; Bera, S.; Karak, S.; Garai, B.; Kharul, U. K.; Banerjee, R. Pore surface engineering in porous, chemically stable covalent organic frameworks for water adsorption. J. Mater. Chem. A 2015, 3, 23664. (23) Held, A.; Pratt, D. W. Ammonia as a hydrogen bond donor and acceptor in the gas phase. Structures of 2-pyridone-NH3 and 2pyridone-(NH3)2 in their S0 and S1 electronic states. J. Am. Chem. Soc. 1993, 115, 9718. (24) Tanner, C.; Manca, C.; Leutwyler, S. Probing the threshold to H atom transfer along a hydrogen-bonded ammonia wire. Science 2003, 302, 1736. (25) Nelson, JR. D. D.; Fraser, G. T.; Klemperer, W. Electric dipole moments of HF−C2H2, HF−C2H4, and HF−C3H6. J. Chem. Phys. 1985, 82, 4483. (26) Taylor, R.; Kennard, O.; Versichel, W. The geometry of the NâHâ -O = C hydrogen bond. 3. Hydrogen-bond distances and angles. Acta Crystallogr., Sect. B: Struct. Sci. 1984, 40, 280. (27) Brown, N. M. D.; Nonhebel, D. C. NMR spectra of intramolecularly hydrogen-bonded compoundsII: Schiff bases of β-diketones and o-hydroxycarbonyl compounds. Tetrahedron 1968, 24, 5655. (28) Golcu, A.; Tumer, M.; Demirelli, H.; Wheatley, R. A. Cd(II) and Cu(II) complexes of polydentate Schiff base ligands: synthesis, characterization, properties and biological activity. Inorg. Chim. Acta 2005, 358, 1785. (29) Meot-Ner, M. The ionic hydrogen bond. Chem. Rev. 2005, 105, 213. (30) Jadhav, S. S.; Kolhe, N. H.; Athare, A. E. Synthesis and characterization of mixed ligand complexes of salicylaldoxime, dimethylglyoxime and benzoin with Mn (II) and their biological activity. Int. J. Pharm. Bio. Sci. 2013, 4 (3), 45−54.

(31) Mehta, R. K.; Gupta, R. K. Iron(II), cobalt(II), nickel(II), copper(II), zinc(II), palladium(II), cadmium(II), and uranium(II) dioxide complexes of tridentate Schiff bases. Indian J. Chem. 1973, 11, 56. (32) Bauschlicher, C. W., Jr; Sodupe, M.; Partridge, H. A theoretical study of the positive and dipositive ions of M(NH3)n and M(H2O)n for M = Mg, Ca, or Sr. J. Chem. Phys. 1992, 96, 4453. (33) Martyna, G. J.; Klein, M. L. Pseudopotential calculation of the electronic states of small metal-ammonia clusters. J. Phys. Chem. 1991, 95, 515. (34) Rogers, R. D.; Jezl, M. L.; Bauer, C. B. Effect of polyethylene glycol on the coordination sphere of strontium in SrCl2 and Sr(NO3)2 complexes. Inorg. Chem. 1994, 33, 5682. (35) Naiini, A. A.; Pinkas, J.; Plass, W.; Young, V. G., Jr.; Verkade, J. G. Triethanolamine complexes of H+, Li+, Na+, Sr2+, and Ba2+ perchlorates. Inorg. Chem. 1994, 33, 2137. (36) Ben, T.; Ren, H.; Ma, S.; Cao, D.; Lan, J.; Jing, X.; Wang, W.; Xu, J.; Deng, F.; Simmons, J. M.; Qiu, S.; Zhu, G. Targeted synthesis of a porous aromatic framework with high stability and exceptionally high surface area. Angew. Chem., Int. Ed. 2009, 48, 9457. (37) Shallenberger, J. R.; Cole, D. A.; Novak, S. W. Characterization of silicon oxynitride thin films by x-ray photoelectron spectroscopy. J. Vac. Sci. Technol., A 1999, 17, 1086. (38) Stevens, J. S.; Byard, S. J.; Seaton, C. C.; Sadiq, G.; Daveya, R. J.; Schroeder, S. L. M. Proton transfer and hydrogen bonding in the organic solid state: a combined XRD/XPS/ssNMR study of 17 organic acid-base complexes. Phys. Chem. Chem. Phys. 2014, 16, 1150. (39) Bisti, F.; Stroppa, A.; Picozzi, S.; Donarelli, M.; Picozzi, S.; Coreno, M. The electronic structure of gas phase croconic acid compared to the condensed phase: more insight into the hydrogen bond interaction. J. Chem. Phys. 2013, 138, 014308. (40) Stevens, J. S.; Byard, S. J.; Seaton, C. C.; Sadiq, G.; Davey, R. J.; Schroeder, S. L. M. Crystallography aided by atomic core-level binding energies: proton transfer versus hydrogen bonding in organic crystal structures. Angew. Chem., Int. Ed. 2011, 50, 9916. (41) Vasquez, R. P. SrCl2 by XPS. Surf. Sci. Spectra 1992, 1, 68. (42) Wang, Z.; Dou, Z.; Cui, Y.; Yang, Y.; Wang, Z.; Qian, G. Nanoassembles constructed from mesoporous silica nanoparticles and surface-coated multilayer polyelectrolytes for controlled drug delivery. Microporous Mesoporous Mater. 2014, 185, 92. (43) Orechovska, J.; Misaelides, P.; Godelitsas, A.; Rajec, P.; KleweNebenius, H.; Noli, F.; Pavlidou, E. Investigation of HEU-type zeolite crystals after interaction with Sr2+ cations in aqueous solution using nuclear and surface analytical techniques. J. Radioanal. Nucl. Chem. 1999, 241, 519. (44) Ong, E. W.; Eckert, J.; Dotson, L. A.; Glaunsinger, W. S. Nature of guest species within alkaline earth-ammonia intercalates of titanium disulfide. Chem. Mater. 1994, 6, 1946. (45) Seredych, M.; Rossin, J. A.; Bandosz, T. J. Changes in graphite oxide texture and chemistry upon oxidation and reduction and their effect on adsorption of ammonia. Carbon 2011, 49, 4392. (46) Micek-Ilnicka, A.; Gil, B.; Lalik, E. Ammonia sorption by Dawson acid studied by IR spectroscopy and microbalance. J. Mol. Struct. 2005, 740, 25. (47) Petit, C.; Seredych, M.; Bandosz, T. Revisiting the chemistry of graphite oxides and its effect on ammonia adsorption. J. Mater. Chem. 2009, 19, 9176. (48) Hadjiivanov, K. FTIR study of CO and NH3 co-adsorption on TiO2 (rutile). Appl. Surf. Sci. 1998, 135, 331. (49) Kung, M. C.; Kung, H. H. IR Studies of NH3, pyridine, CO, and NO adsorbed on transition metal oxides. Catal. Rev.: Sci. Eng. 1985, 27, 425.

G

DOI: 10.1021/acscentsci.8b00232 ACS Cent. Sci. XXXX, XXX, XXX−XXX