Organic Framework Synthesis via Retrosynthetic ... - ACS Publications

Dec 12, 2018 - Department of Materials Science and Engineering, Texas A&M ... MOF with high guest capacity, highlighting the integrated cooperative be...
0 downloads 0 Views 7MB 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

http://pubs.acs.org/journal/acscii

Uncovering Two Principles of Multivariate Hierarchical Metal− Organic Framework Synthesis via Retrosynthetic Design Liang Feng,† Shuai Yuan,† Jia-Luo Li,† Kun-Yu Wang,† Gregory S. Day,† Peng Zhang,† Ying Wang,*,†,‡,∥ and Hong-Cai Zhou*,†,§ †

Department of Chemistry, Texas A&M University, College Station, Texas 77843-3255, United States College of Chemistry, Tianjin Normal University, Tianjin, 300387, China § Department of Materials Science and Engineering, Texas A&M University, College Station, Texas 77843-3003, United States ∥ Key Laboratory of Advanced Energy Materials Chemistry (Ministry of Education), Nankai University, Tianjin 300071, China

Downloaded via UNIV OF SOUTH DAKOTA on December 16, 2018 at 08:25:21 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.



S Supporting Information *

ABSTRACT: Multivariate (MTV) hierarchical metal−organic frameworks (MOFs), which contain multiple regions arranged in ordered structures, show promise for applications such as gas separation, size-selective catalysis, and controlled drug delivery. However, the complexity of these hierarchical MOFs is limited by a lack of control during framework assembly. Herein, we report the controlled generation of hierarchical MOF-on-MOF structural formation under the guidance of two design principles, surface functionalization and retrosynthetic techniques for stability control. Accordingly, the tunability of spatial distributions, compositions, and crystal sizes has been achieved in these hierarchical systems. The resulting MOF-on-MOF hierarchical structures represent a unique crystalline porous material which contains a controllable distribution of functional groups and metal clusters that are associated together within a framework composite. This general synthetic approach not only expands the scope and tunability of the traditional MTV strategy to multicomponent materials, but also offers a facile route to introduce variants and sequences to sophisticated three-dimensional hierarchical and cooperative systems. As a proof of concept, the photothermal effects of a porphyrinic core-MOF are exploited to trigger the controlled guest release from a shellMOF with high guest capacity, highlighting the integrated cooperative behaviors in multivariate hierarchical systems.



INTRODUCTION The precise location engineering of building units in proteins and nucleic acids is essential for chemical processes in cells such as recognition, biocatalysis, and information storage. These well-defined sequences and distributions have inspired researchers to develop tailored architectures with controllable heterogeneity in polymers, nanomaterials, and porous materials.1−3 For example, recent work on bimetallic nanoparticles synthesized through surface inorganometallic chemistry demonstrated enhanced catalytic performance due to the synergistic effect between constituent metals.4 Precise molecular encoding on multicomponent polymers has also had a substantial influence on their macroscopic properties.5−7 Multivariate (MTV) metal−organic frameworks (MOFs) constructed from multiple metal clusters and linkers promise fascinating advances in fields such as gas separation, sizeselective catalysis, and controlled drug delivery.8−17 Among these MOFs, MTV-MOF-5, developed by Yaghi and co© XXXX American Chemical Society

workers, contains up to eight different functionalities in a disordered arrangement in an overall ordered backbone framework.18 Furthermore, we have demonstrated that the installation of linear linkers and inorganic clusters in coordinatively unsaturated MOFs can postsynthetically arrange functionalities into predicted positions.19−21 Alternatively, the construction of hierarchical MTV-MOFs from multiple components is considered a viable pathway for achieving advanced applications that require sophisticated architectures.22−28 However, the growing gap between the design and synthesis of hierarchical MOFs has become a critical limitation mainly because lattice matching (similar crystallographic parameters) has typically been required for the epitaxial growth method (EGM) necessary to achieve hierarchical MOF-on-MOF hybrids growth. Yet, as most Received: October 6, 2018

A

DOI: 10.1021/acscentsci.8b00722 ACS Cent. Sci. XXXX, XXX, XXX−XXX

Research Article

ACS Central Science

Scheme 1. Preparation of Multivariate Hierarchical Metal−Organic Frameworks through Retrosynthetic Designa

a

(a) Direct reaction between PCN-222 and MOF-5 precursor leads to the homogeneous nucleation and growth of MOF-5 as a separate phase; (b) overcoming the energy barrier of heterogeneous nucleation of MOF-5 by surface functionalization; (c) retrosynthetic analysis of bimetallic MOFs starting from monometallic MOFs exhibiting surface defects. This strategy can be utilized to guide the stepwise synthesis of hierarchical PCN222@MOF-5 composites. Note that the ball-and-stick model of MOF-5 refers to the whole crystal lattice. The scale of PCN-222 is narrowed to better present the apportionment of PCN-222 inside MOF-5.

designed MOF structures with intrinsic defects, can lead to sophisticated structures by utilizing kinetic control and postsynthetic modifications.32,33 Similarly, the kinetically controlled synthesis of hierarchical MOFs is designed to proceed in the presence of the coordination processes occurring during surface functionalization. The hierarchical structure forms after the self-assembly of the shell MOF outside the core MOFs. The whole process is controlled by kinetics so that different components can be precisely placed outside the original core MOF crystal lattice while maintaining the structural integrity of the core MOF.

MOFs have distinct crystallographic parameters, EGM has been severely limited in applicability. A recently developed method by Kitagawa and co-workers involves the introduction of a polymer to connect the two different MOF phases with varying crystallographic parameters into one unified composite.29 However, this method cannot achieve full coverage of the core-MOFs with the shell-MOFs necessary for size selective applications. Besides, the introduction of the polymer into the interface between the lattice mismatched core and shell-MOFs can potentially undermine the intricate MOF− MOF interface, which often includes defect sites and coordinatively unsaturated metal sites. Recently, our group reported a one-step synthesis of hybrid core−shell MOFs with mismatched lattices under the guidance of nucleation kinetic analysis.30 Yet, the versatility of this one-pot nucleation kinetic control strategy is limited because it is challenging to generate clusters with different metal species under one-pot synthetic conditions. Thus, the development of a novel and facile method for the rational design and synthesis of hierarchical MOFs without additives, while still being able to overcome the restrictions of the lattice matching rule, is quite urgent. In this work, we report two principles for preparing hierarchical MOFs with mismatched lattices through retrosynthetic design: surface functionalization and retrosynthetic stability considerations.19,31 Retrosynthetic analysis has been mainly utilized in the synthesis of organic molecules based on covalent bonds by transformation of a target molecule into simpler precursors and sequential assembly through a set of known chemical reactions. We believe that the conceptual scope of retrosynthetic analysis can be extended to coordination bond based complicated systems, for example, MTV hierarchical MOFs (Figure S18). Retrosynthetically assembled metal precursors and ligands, placed inside the



RESULTS AND DISCUSSION

Principle 1: Surface Functionalization. A zirconiumMOF, PCN-222 (also known as MOF-545),34,35 was used as a core scaffold for the growth of a MOF shell because the high chemical stability of PCN-222 ensures its structural integrity under postsynthetic modification conditions. On the basis of Pearson’s hard/soft acid/base (HSAB) principle, MOFs based on different metal clusters should have distinct stabilities determined by the MO bond strength.36 Generally, the harsh solvothermal synthetic conditions of Zr-MOFs are not suitable for Zn-MOFs. Zhong and co-workers reported that Zn-MOFs gradually dissolve and become metal-linker complexes under the synthetic condition of UiO-66 (Zr).37 In contrast, Zr-MOFs are robust frameworks that should be able to survive under the synthetic conditions of Zn-MOFs. However, due to the increased energy barrier caused by the mismatched lattice parameters, MOF-538 tends to grow separately when its precursor components are directly mixed with PCN-222. The heterogeneous nucleation of the Zn-MOF B

DOI: 10.1021/acscentsci.8b00722 ACS Cent. Sci. XXXX, XXX, XXX−XXX

Research Article

ACS Central Science

S7). The morphology of PCN-222 was well maintained in the hierarchical structure. Powder X-ray diffraction (PXRD) patterns of the hierarchical MOFs demonstrated the wellmaintained crystallinity of both core and shell MOFs (Figure 1e). Scanning electron microscopy/energy dispersive X-ray spectroscopy (SEM-EDS) and confocal laser scanning microscopy indicated that Zr was well dispersed in the lattice of the Zn-MOFs. The Zr/Zn ratio of both the internal and external structures of PCN-222@MOF-5 calculated from EDS mapping were consistent with the results from ICP-MS data of the digested MOF samples (Figure S6). The PCN-222 to MOF-5 ratio could be well tuned in the hierarchical PCN-222@MOF-5 composites by adjusting the feed ratios, producing PCN-222@MOF-5-(R%), where R% is the mole percentage of PCN-222 as determined by 1H NMR (Figures S12−S14, Table S1). The N2 adsorption isotherms of the hierarchical PCN-222@MOF-5 were measured, showing hierarchical porosity as expected (Figure 1d). The total N2 uptakes were close to those of the shell MOFs, while a noticeable hysteresis loop near P/P0 = 0.3 confirmed the presence of the core MOFs. The pore size distributions of hierarchical MOFs were calculated from the N2 adsorption isotherms using a density functional theory (DFT) model, showing the characteristic 32 Å peak of PCN-222 (Figure S55). In contrast to our previously reported top-down approaches for hierarchical structures, linker labilization39 and thermolysis,22 this retrosynthetic design showcases an alternative bottom-up approach for the design of sophisticated hierarchical structures. To demonstrate the importance of surface functionalization, a control experiment, inspired by a cooperative template system reported by us previously, was designed.40 The cooperative template system utilized a chelating agent (citric acid, CA) as an interface between the surfactant and MOF surfaces. Here, we functionalized the surface of PCN-222 with the carboxylate groups of CA, which can also generate the hierarchical MOF composites (Figures S10−S11). These results indicate that surface functionalization could help to connect the two different MOFs with totally different lattice parameters by lowering the surface energy. Additionally, it should be noted that the slow release kinetics of linkers on the surface of the MOFs can also contribute to the gradual formation of the MOF-on-MOF structures. Recently, the Matzger group reported the slow diffusion phenomenon during the linker exchange process, which originated from hydrogen bonding between the ligands.41 Similarly, the diffusion of linkers from MOF pores to the interface can also be limited under comparable conditions (Figure S16). Principle 2: Retrosynthetic Stability Consideration. To examine whether kinetically guided synthesis could function as a general strategy, we extended this approach to other common MOF systems to develop sophisticated hierarchical MOF composites (Figure 2a). We chose M(II) based carboxylate MOFs as the shell MOFs, for example, the MOF-5 (Zn) series, MOF-177 (Zn), HKUST-1 (Cu), and MOF-1114 (Yb) due to their mild synthetic conditions, relative poor stability in acidic environments, and relatively large crystal sizes (Figures S19−S23). Meanwhile, stable MOFs, including UiO-66 (Zr), UiO-67 (Zr), PCN-160 (Zr), MOF-808 (Zr), PCN-222 (Zr), PCN-250 (Fe), and MIL-125 (Ti), were selected as core MOFs because of their high chemical stability (Figures S24−S27). On the basis of the principles discussed above, we were able to construct

is not favored, leading to the formation of Zn- and Zr-MOF mixtures (Scheme 1a). To lower the energy barrier of heterogeneous nucleation during hierarchical MOF formation, the surfaces of the core MOFs were functionalized by pretreating with excess linker under solvothermal conditions (Scheme 1b−c). This allowed for the coordination of the linkers on the surface of the core MOFs and also resulted in the capture of preloaded linkers inside the MOF pores. Further addition of metal precursors to the system allowed for the gradual formation of shell MOFs on the surface of the core MOFs (Figure S16). The crystals of PCN-222 were further treated with an excess of benzene-1,4dicarboxylic acid (BDC) as a dimethylformamide (DEF) solution for 24 h, followed by a DEF solution of Zn(NO3)2 for another 24 h. As a result, a hierarchical MOF-on-MOF composite, PCN-222@MOF-5, was obtained. The formation of hierarchical MOFs could be clearly distinguished from optical microscopy images of the respective crystals (Figure 1a). The core MOF, porphyrinic PCN-222

Figure 1. Tunable compositions in hierarchical PCN-222@MOF-5. (a) Models and optical images showing controlled incorporation of PCN-222 inside MOF-5, from 0% to 14%. (b−c) SEM images of the surface (b) and interior (c) of PCN-222@MOF-5-(8%). The internal areas of PCN-222@MOF-5-(8%) were accessed by physically crushing samples. (d−e) N2 sorption isotherms and PXRD patterns of PCN-222@MOF-5-(R%, R = 1, 8, and 14).

constructed from the four connected linker tetrakis(4carboxyphenyl)porphyrin (TCPP), exhibited a purple color and a needle-like shape, whereas the shell MOF-5 exhibited colorless cubic features. Scanning electron microscopy (SEM) revealed that the needle-like PCN-222 crystals were well dispersed both inside and outside MOF-5 (Figures 1b−c and C

DOI: 10.1021/acscentsci.8b00722 ACS Cent. Sci. XXXX, XXX, XXX−XXX

Research Article

ACS Central Science

Figure 2. Versatility of kinetically guided retrosynthesis of hierarchical MOFs. (a) Sequence-defined combinations of clusters and linkers into hierarchical MOFs with controllable compositions and distributions. More stable MOFs with high valence metals (M4+, M3+) should be designed as core MOFs, while less stable MOFs with low valence metals (M+, M2+) would be assigned as shell MOFs. (b−d, f−j) Optical images of hierarchical MOFs with sequence-defined combinations guided by retrosynthetic stability consideration. Scale bar is 100 μm in all optical images. (e) Unlimited combination of multivariate hierarchical MOFs that can be synthesized under the guidance of the two principles.

multicomponent hierarchical MOFs with in a large number of combinations (Figures 2e and S17−S18). The presence of both core MOFs and shell MOFs was confirmed by optical imaging and PXRD patterns (Figures 2b−d,f−j and S28−S51). The porosity and thermal stability of these hierarchical MOFs are summarized in Figures S53−S59 and Table S3. Our retrosynthetic design exhibits a general and powerful tool to control the distribution inside multivariate hierarchical MOFs. Interestingly, by controlling the apportionment of core MOFs dispersed in shell MOFs, we were able to get three types of hierarchical MOF-on-MOF structures: well-mixed, center-concentrated, and half−half asymmetric distribution (Figure 3). Through centrifugal precipitation, core MOF (PCN-222) powders could be well stabilized at the bottom of the reaction vessels, allowing for the subsequent anisotropic growth of MOF-5 (Figure 3c). After the heterogeneous nucleation and growth, asymmetric apportionment (similar to Janus particles) were generated as a result (Figure 3d). In contrast, dispersing PCN-222 nanocrystals in a MOF-5 precursor solution through ultrasonic mixing followed by the crystallization process could lead to the formation of wellmixed “solid solutions” (PCN-222@MOF-5, Figure 3a−b). Applying higher concentrations of MOF-5 precursors to the synthesis of hierarchical PCN-222@MOF-5 would further promote the formation of center-concentrated core−shell structures, as indicated by optical imaging (Figure 3a−b). In this case, the cubic core was observed as the well mixed PCN222@MOF-5, which facilitated the further growth of lattice matched shell MOF-5. Additionally, if both the core and shell MOFs have comparable sizes, core−shell structures, instead of solid solutions, would form directly. For instance, as shown in Figure S48, the stepwise evolution of the single-crystalline core−shell HKUST-1@MOF-5 was observed.

Figure 3. Tuning spatial distributions in hierarchical PCN-222@ MOF-5. (a−b) Evolution of well-mixed and center-concentrated PCN-222@MOF-5 hierarchical MOFs. (c−d) Preparation of an asymmetric dispersion of PCN-222 in MOF-5 shell (similar to Janus particles). Scale bar is 100 μm in inset images.

By utilizing the two principles described herein, numerous hierarchical MTV-MOFs could be designed and synthesized, with even multilayer MTV MOFs being formed, expanding the scope of traditional MTV strategy for multicomponent materials. Among the prepared hierarchical MOFs, PCN222@ZIF-8 shows extraordinary stability in water, maintaining structural integrity in boiling water for more than 24 h (Figure D

DOI: 10.1021/acscentsci.8b00722 ACS Cent. Sci. XXXX, XXX, XXX−XXX

Research Article

ACS Central Science S51). Compared with the traditional one-pot introduction of functionality inside of MOFs, this retrosynthetic synthesis of PCN-222@ZIF-8 provides a method for introducing linkers with various functional groups and connectivities without interfering with the growth process. This heterogeneous growth (without TCPP competition) helped us obtain highly crystalline MOFs and overcome the impurity problems associated with the homogeneous nucleation process (resulting from TCPP competition). Porphyrins or metalloporphyrins show wide applications in areas such as catalysis, photochemistry, and biological applications; therefore, the efficient immobilization of porphyrins in porous supports with controllable pore environment is highly desired for targeted applications. In contrast, we have shown that TCPP@ZIF-8, synthesized by a one-pot solvothermal reaction between TCPP, Zn2+, 2-methylimidazole (mIM) and DMF, shows a relative low level of crystallinity for ZIF-8, possibly due to the competitive coordination between TCPP and mIM under homogeneous conditions (Figure S52). However, the introduction of heterogeneous PCN-222 as seeds can effectively avoid the competition, generating highly crystalline ZIF-8 with porphyrin units well dispersed inside of ZIF-8. In addition, TCPP is more firmly immobilized in PCN-222@ZIF8 than in TCPP@ZIF-8 when exposed to harsh conditions due to the presence of the more robust coordination bonds between TCPP and Zr clusters. Size-Selective Catalytic Platform. We have shown that a size-selective biomimetic catalytic system PCN-222(Fe)@ZIF8 could be constructed by the strategy presented here. The FeTCPP unit in the core-MOF, PCN-222(Fe), provides the catalytic active site while the shell-MOF, ZIF-8, controls the size selectivity of substrates (Figure 4a). To evaluate the accessibility of FeTCPP in hierarchical MOFs, the oxidation of two model molecules, o-phenylenediamine (o-PDA) and 2,2′azino-bis(3-ethylbenzothiazoline-6-sulfonic acid) (ABTS), was performed (Figures S64−S65). The reactions were conducted directly in 3 mL cuvettes. Before incorporation, MOF samples were activated by solvent exchange with DMF and dichloromethane three times and further dried overnight at 70 °C under a vacuum. The substrate (1 mL) and 1 mg of MOF catalysts were added to buffer followed with the addition of H2O2 (1 mL). After shaking for a short while, the cuvette was put in the UV sample holder as quickly as possible. The data were collected using a kinetics mode, and the plots of absorbance versus time were obtained. As indicated by Figure 4b and Table S4, PCN-222(Fe) shows slightly higher kcat, Km, and Vmax values compared to PCN-222(Fe)@ZIF-8 during the catalytic transformation of o-PDA. The molecular size of oPDA (0.5 × 0.5 nm) is relatively small, allowing it to diffuse through the ZIF-8 window and reach the active sites of the encapsulated PCN-222(Fe), leading to considerable activity compared to PCN-222(Fe) without shell protection. However, due to the larger size of ABTS (0.7 × 1.6 nm), the diffusion of ABTS through ZIF-8 is largely restricted, which in turn results in a dramatically decreased kcat and Vmax compared to the free PCN-222(Fe), highlighting the beneficial size-selective effect of hierarchical core−shell structure on the catalytic performance. Phototriggered Guest Release. Coupling functionalities into specific isolated layers should help generate individual workshops for a single task, while integrating these workshops into a hierarchical system would allow for the production of multitask targets. As a proof of concept, we design a

Figure 4. (a) Tunability of the pore environment and guest penetration behavior in hierarchical MOFs for tailored applications. (b) Relative activity of PCN-222(Fe) and hierarchical PCN-222(Fe) @ZIF-8 for the oxidation of o-phenylenediamine (o-PDA) and ABTS.

hierarchical system with the ability to engage in phototriggered guest release. Interestingly, it has recently been shown that porphyrin-based MOFs such as PCN-223 have a photothermal effect when irradiated under visible light (Figure 5b), while nonporphyrin-based MOFs, such as ZIF-8 and MOF-177, do not exhibit this response (Figure 5a).42 It should be noted that PCN-223, which only has micropores and a relatively low surface area, would not be an ideal platform for guest uptake, while ZIF-8 and MOF-177, both of which possess large cages, are capable of capturing large amounts of guests within their pores (Table S6). Therefore, the design of the shell-MOFs (ZIF-8 or MOF-177) takes advantage of the high porosity and guest storage capacity of the shell MOFs, while the core-MOFs (such as PCN-223) are utilized as a photoresponsive material to trigger the photothermal desorption of the guests. The photothermal effects of MOFs with or without porphyrinic units were first examined by immersing 20 mg of the MOF samples in 10 mL of solvent under the irradiation of LED lamps (λ = 410 nm, 150 W). Notably, after irradiation for 30 min, the system temperature spontaneously increased from 22.0 to 28.6 °C for PCN-223, while the resulting temperature (22.1 °C) of the system in the absence of catalyst or the presence of only ZIF-8 showed negligible changes (Table S5). To study this photoresponsive behavior, hierarchical PCN223@MOF-177 and PCN-223@ZIF-8 were then tested for the release of 4-nitrophenol and anthracene. The two model guests were selected due to their suitable sizes and facile detection approach via UV−vis spectroscopy. The guest incorporation within the pores of MOF-177, ZIF-8, PCN-223, and their E

DOI: 10.1021/acscentsci.8b00722 ACS Cent. Sci. XXXX, XXX, XXX−XXX

Research Article

ACS Central Science

centrifuged, and the UV−vis spectra of the supernatants with and without lamp irradiation were recorded and compared to each other. The release amount can be calculated from the standard curve (anthracene at λ = 355 nm, 4-nitrophenol at λ = 282 nm). As indicated by Figure 6, UV intensities recorded

Figure 6. Phototriggered release of 4-nitrophenol as a function of time with and without lamp irradiation determined by UV−vis spectra using the kinetic mode. Hierarchical PCN-223@ZIF-8 and PCN223@MOF-177 can achieve more efficient phototriggered release than its single component or physical mixtures.

Figure 5. Phototriggered guest release by multivariate hierarchical MOFs. (a) Shell MOFs without photoresponsive units cannot release guests upon light exposure; (b) core MOFs with photothermal effects lack the controlled release of guests and have overall guest capacity; (c) physical mixtures of (a, b) result in low photoresponsive efficiency due to the heat loss to surrounding solvents; (d) multivariate hierarchical MOFs with well designed apportionments can achieve efficient phototriggered release with high guest capacity and controllable release.

in the PCN-223@MOF-177 system without light exposure showed almost no change, indicating the absence of 4nitrophenol release. This can be explained by the strong interaction between 4-nitrophenol and the host framework, which dramatically slows down the release process. Remarkably, for the irradiated PCN-223@MOF-177, an efficient phototriggered release of 4-nitrophenol was observed. The amount of released 4-nitrophenol was increased as the time period extended, demonstrating the effective heat transfer from core PCN-223 to the surrounding guest molecules under exposure from the 410 nm light. Similarly, PCN-223@ZIF-8 also exhibited this phototriggered release behavior, but with a slightly slower rate due to the lower porosity and smaller pore window size of ZIF-8 compared with MOF-177. In contrast, MOF-177 or ZIF-8, in the absence of core PCN-223, did not exhibit phototriggered release because they were unable to convert optical energy to heat, which is required for targeted release of guest molecules. Reference experiments using physical mixtures of PCN-223/MOF-177 were performed with a much lower releasing rate, possibly because a large portion of heat was lost during the transfer from PCN-223 to MOF-177 (Figure 5c). Note that the well-mixed nature of the hierarchical PCN-223@MOF-177 prevented unnecessary heat loss to solvents and ensured the effective guest desorption by localized heat (Figure 5d). Therefore, the hierarchical PCN223@MOF-177 can achieve more efficient photoresponsive releasing behavior compared to its constituent components or even their mixtures. Additionally, this phototriggered behavior can also be observed by utilizing anthracene as a model molecule, released from hierarchical PCN-223@ZIF-8 (Figures S68−S69).

hierarchical structures was first studied. Before incorporation, MOF samples were activated by solvent exchange with DMF and dichloromethane three times and further dried overnight at 70 °C under a vacuum. The guest loadings were then performed by soaking 10 mg of activated MOFs (PCN-223, ZIF-8, MOF-177, and the hierarchical PCN-223@MOF-177, PCN-223@ZIF-8) in a 20 mM cyclohexane solution of anthracene (or 20 mM 4-nitrophenol, 2 mL) in a thermomixer at room temperature for 24 h. After guest insertion, the MOF samples were washed with fresh cyclohexane three times to remove the guest molecules on the surface of the crystals, which was confirmed by UV−vis spectroscopy. The loading of anthracene (or 4-nitrophenol) in various MOF samples was calculated by the comparison of the UV−vis absorbance (anthracene at λ = 355 nm, 4-nitrophenol at λ = 282 nm). The capacity of the guest loading was summarized in Table S6. The capacity of the 4-nitrophenol in corresponding frameworks increases in the order of PCN-223 < PCN-223@ZIF-8 < ZIF-8 < PCN-223@MOF-177 < MOF-177, which is accompanied by an increase in porosity and pore volume. Release experiments were conducted by using an LED light source with an emission wavelength at 410 nm, and the release amounts of the guest molecules were further analyzed by UV− vis spectroscopy. Ten milligrams of guest loaded MOFs was dispersed into 1 mL of cyclohexane. The sample was then directly irradiated with the LED lamp (150 W, λ = 410 nm) for various periods (0−5 h). After each experiment, samples were F

DOI: 10.1021/acscentsci.8b00722 ACS Cent. Sci. XXXX, XXX, XXX−XXX

Research Article

ACS Central Science



CONCLUSIONS Altogether, these results exemplify the capability of retrosynthetically guided preparation of multivariate hierarchical systems with a tailored pore environment and cooperative behaviors, giving an unparalleled level of control over guest storage and transportation. This kinetically guided synthesis of hierarchical MOFs offers an ideal platform for the systematic tuning of pore environments and guest penetration behavior by the judicious choosing of clusters and linkers on core and shell MOFs. The current results already show the benefit of systems such as these for substrate selectivity processes. The incorporation of size selective components as a shell over a catalytically active porphyrinic MOF core has already shown applicability for the size selective oxidation of organic substrates. These results show the possibility of these core− shell structures and provide a general method that should allow for the controlled growth of tailored catalytic and capture materials. Potential applications include using a core MOF designed as a mesoporous MOF with preloaded drugs, while the shell MOF could be constructed with multiply functional groups that engage in favorable interactions with either the guest molecules or particular cell environments. This would allow for the controlled release of drug molecules to areas of interest in the body. In addition, the heterogeneous nucleation of the shell MOF on the core MOF can allow for the controlled growth of materials that would be incompatible in a one-pot systems, such as growing an acidic MOF onto the surface of a basic MOF.



U.S. Department of Energy Office of Fossil Energy, National Energy Technology Laboratory (DE-FE0026472), and National Science Foundation Small Business Innovation Research (NSF-SBIR) under Grant No. (1632486), National Natural Science Foundation of China (21471113), Training Program of Outstanding Youth Innovation Team of Tianjin Normal University, National Top-Level Talent Training Program of Tianjin Normal University, Training Program of Young and middle-aged Innovative Talent and Backbone, Tianjin “131” Innovative Talent, 111 project (B12015), and Opening Fund of Key Laboratory of Advanced Energy Materials Chemistry (Ministry of Education), Nankai University. Use of the TAMU Materials Characterization Facility is acknowledged.



(1) Ida, S.; Ouchi, M.; Sawamoto, M. Template-assisted selective radical addition toward sequence-regulated polymerization: lariat capture of target monomer by template initiator. J. Am. Chem. Soc. 2010, 132 (42), 14748−14750. (2) Kametani, Y.; Sawamoto, M.; Ouchi, M. Control of the alternating sequence for N-isopropylacrylamide (NIPAM) and methacrylic acid units in a copolymer by cyclopolymerization and transformation of the cyclopendant group. Angew. Chem., Int. Ed. 2018, 57 (34), 10905−10909. (3) Khaletskaya, K.; Reboul, J.; Meilikhov, M.; Nakahama, M.; Diring, S.; Tsujimoto, M.; Isoda, S.; Kim, F.; Kamei, K. I.; Fischer, R. A.; Kitagawa, S.; Furukawa, S. Integration of porous coordination polymers and gold nanorods into core-shell mesoscopic composites toward light-induced molecular release. J. Am. Chem. Soc. 2013, 135 (30), 10998−11005. (4) Ding, K.; Cullen, D. A.; Zhang, L.; Cao, Z.; Roy, A. D.; Ivanov, I. N.; Cao, D. A general synthesis approach for supported bimetallic nanoparticles via surface inorganometallic chemistry. Science 2018, 362, 560−564. (5) Gutekunst, W. R.; Hawker, C. J. A general approach to sequencecontrolled polymers using macrocyclic ring opening metathesis polymerization. J. Am. Chem. Soc. 2015, 137 (25), 8038−8041. (6) Pfeifer, S.; Lutz, J. F. A facile procedure for controlling monomer sequence distribution in radical chain polymerizations. J. Am. Chem. Soc. 2007, 129 (31), 9542−9543. (7) Lutz, J. F.; Ouchi, M.; Liu, D. R.; Sawamoto, M. Sequencecontrolled polymers. Science 2013, 341 (6146), 1238149. (8) Popp, T. M. O.; Yaghi, O. M. Sequence-dependent materials. Acc. Chem. Res. 2017, 50 (3), 532−534. (9) Furukawa, H.; Muller, U.; Yaghi, O. M. “Heterogeneity within order” in metal-organic frameworks. Angew. Chem., Int. Ed. 2015, 54 (11), 3417−3430. (10) Zhou, H. C.; Long, J. R.; Yaghi, O. M. Introduction to metalorganic frameworks. Chem. Rev. 2012, 112 (2), 673−674. (11) Dong, Z. Y.; Sun, Y. Z. S.; Chu, J.; Zhang, X. Z.; Deng, H. X. Multivariate metal-organic frameworks for dialing-in the binding and programming the release of drug molecules. J. Am. Chem. Soc. 2017, 139 (40), 14209−14216. (12) Zhou, H. C.; Kitagawa, S. Metal-organic frameworks (MOFs). Chem. Soc. Rev. 2014, 43 (16), 5415−5418. (13) Li, J. R.; Sculley, J.; Zhou, H. C. Metal-organic frameworks for separations. Chem. Rev. 2012, 112 (2), 869−932. (14) Lee, J.; Farha, O. K.; Roberts, J.; Scheidt, K. A.; Nguyen, S. T.; Hupp, J. T. Metal-organic framework materials as catalysts. Chem. Soc. Rev. 2009, 38 (5), 1450−1459. (15) Kirchon, A.; Feng, L.; Drake, H. F.; Joseph, E. A.; Zhou, H. C. From fundamentals to applications: a toolbox for robust and multifunctional MOF materials. Chem. Soc. Rev. 2018, 47, 8611− 8638. (16) Helal, A.; Yamani, Z. H.; Cordova, K. E.; Yaghi, O. M. Multivariate metal-organic frameworks. Natl. Sci. Rev. 2017, 4 (3), 296−298.

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acscentsci.8b00722. Materials and methods, synthesis and characterization, and catalytic experiment details (PDF)



REFERENCES

AUTHOR INFORMATION

Corresponding Authors

*(Y.W.) E-mail: [email protected]. *(H.-C.Z.) E-mail: [email protected]. ORCID

Liang Feng: 0000-0001-9693-1595 Hong-Cai Zhou: 0000-0002-9029-3788 Notes

No unexpected or unusually high safety hazards were encountered. The authors declare no competing financial interest.



ACKNOWLEDGMENTS The gas adsorption−desorption studies of this research were supported as part of the Center for Gas Separations Relevant to Clean Energy Technologies, an Energy Frontier Research Center funded by the U.S. Department of Energy, Office of Science, Basic Energy Sciences under Award Number DESC0001015. The synthesis and PXRD, TGA, and SEM characterization and analysis were funded by the Robert A. Welch Foundation through a Welch Endowed Chair to H.J.Z. (A-0030). The catalysis work was funded by the Qatar National Research Fund under Award Number NPRP9-377-1080. The authors also acknowledge the financial supports of G

DOI: 10.1021/acscentsci.8b00722 ACS Cent. Sci. XXXX, XXX, XXX−XXX

Research Article

ACS Central Science (17) Qin, J. S.; Yuan, S.; Wang, Q.; Alsalme, A.; Zhou, H. C. Mixedlinker strategy for the construction of multifunctional metal-organic frameworks. J. Mater. Chem. A 2017, 5 (9), 4280−4291. (18) Deng, H. X.; Doonan, C. J.; Furukawa, H.; Ferreira, R. B.; Towne, J.; Knobler, C. B.; Wang, B.; Yaghi, O. M. Multiple functional groups of varying ratios in metal-organic frameworks. Science 2010, 327 (5967), 846−850. (19) Yuan, S.; Qin, J. S.; Li, J. L.; Huang, L.; Feng, L.; Fang, Y.; Lollar, C.; Pang, J. D.; Zhang, L. L.; Sun, D.; Alsalme, A.; Cagin, T.; Zhou, H. C. Retrosynthesis of multi-component metal-organic frameworks. Nat. Commun. 2018, 9, 808. (20) Yuan, S.; Chen, Y.-P.; Qin, J.-S.; Lu, W.; Zou, L.; Zhang, Q.; Wang, X.; Sun, X.; Zhou, H.-C. Linker installation: engineering pore environment with precisely placed functionalities in zirconium MOFs. J. Am. Chem. Soc. 2016, 138 (28), 8912−8919. (21) Yuan, S.; Lu, W.; Chen, Y.-P.; Zhang, Q.; Liu, T.-F.; Feng, D.; Wang, X.; Qin, J.; Zhou, H.-C. Sequential linker installation: precise placement of functional groups in multivariate metal−organic frameworks. J. Am. Chem. Soc. 2015, 137 (9), 3177−3180. (22) Feng, L.; Yuan, S.; Zhang, L. L.; Tan, K.; Li, J. L.; Kirchon, A.; Liu, L. M.; Zhang, P.; Han, Y.; Chabal, Y. J.; Zhou, H. C. Creating hierarchical pores by controlled linker thermolysis in multivariate metal-organic frameworks. J. Am. Chem. Soc. 2018, 140 (6), 2363− 2372. (23) Shen, K.; Zhang, L.; Chen, X. D.; Liu, L. M.; Zhang, D. L.; Han, Y.; Chen, J. Y.; Long, J. L.; Luque, R.; Li, Y. W.; Chen, B. L. Ordered macro-microporous metal-organic framework single crystals. Science 2018, 359 (6372), 206−210. (24) Furukawa, S.; Hirai, K.; Nakagawa, K.; Takashima, Y.; Matsuda, R.; Tsuruoka, T.; Kondo, M.; Haruki, R.; Tanaka, D.; Sakamoto, H.; Shimomura, S.; Sakata, O.; Kitagawa, S. Heterogeneously hybridized porous coordination polymer crystals: fabrication of heterometallic core-shell single crystals with an in-plane rotational epitaxial relationship. Angew. Chem., Int. Ed. 2009, 48 (10), 1766−1770. (25) Li, T.; Sullivan, J. E.; Rosi, N. L. Design and preparation of a core-shell metal-organic framework for selective CO2 capture. J. Am. Chem. Soc. 2013, 135 (27), 9984−9987. (26) Choi, S.; Kim, T.; Ji, H.; Lee, H. J.; Oh, M. Isotropic and anisotropic growth of metal-organic framework (MOF) on MOF: logical inference on MOF structure based on growth behavior and morphological feature. J. Am. Chem. Soc. 2016, 138 (43), 14434− 14440. (27) Zhao, M. T.; Yuan, K.; Wang, Y.; Li, G. D.; Guo, J.; Gu, L.; Hu, W. P.; Zhao, H. J.; Tang, Z. Y. Metal-organic frameworks as selectivity regulators for hydrogenation reactions. Nature 2016, 539 (7627), 76− 80. (28) Hirai, K.; Furukawa, S.; Kondo, M.; Uehara, H.; Sakata, O.; Kitagawa, S. Sequential functionalization of porous coordination polymer crystals. Angew. Chem., Int. Ed. 2011, 50 (35), 8057−8061. (29) Gu, Y. F.; Wu, Y. N.; Li, L. C.; Chen, W.; Li, F. T.; Kitagawa, S. Controllable modular growth of hierarchical MOF-on-MOF architectures. Angew. Chem., Int. Ed. 2017, 56 (49), 15658−15662. (30) Yang, X. Y.; Yuan, S.; Zou, L. F.; Drake, H.; Zhang, Y. M.; Qin, J. S.; Alsalme, A.; Zhou, H. C. One-step synthesis of hybrid core-shell metal-organic frameworks. Angew. Chem., Int. Ed. 2018, 57 (15), 3927−3932. (31) Corey, E. J. The logic of chemical synthesis - multistep synthesis of complex carbogenic molecules. Angew. Chem., Int. Ed. Engl. 1991, 30 (5), 455−465. (32) Evans, J. D.; Sumby, C. J.; Doonan, C. J. Post-synthetic metalation of metal-organic frameworks. Chem. Soc. Rev. 2014, 43 (16), 5933−5951. (33) Cohen, S. M. Postsynthetic methods for the functionalization of metal-organic frameworks. Chem. Rev. 2012, 112 (2), 970−1000. (34) Feng, D. W.; Gu, Z. Y.; Li, J. R.; Jiang, H. L.; Wei, Z. W.; Zhou, H. C. Zirconium-metalloporphyrin PCN-222: mesoporous metalorganic frameworks with ultrahigh stability as biomimetic catalysts. Angew. Chem., Int. Ed. 2012, 51 (41), 10307−10310.

(35) Morris, W.; Volosskiy, B.; Demir, S.; Gandara, F.; McGrier, P. L.; Furukawa, H.; Cascio, D.; Stoddart, J. F.; Yaghi, O. M. Synthesis, structure, and metalation of two new highly porous zirconium metalorganic frameworks. Inorg. Chem. 2012, 51 (12), 6443−6445. (36) Yuan, S.; Feng, L.; Wang, K.; Pang, J.; Bosch, M.; Lollar, C.; Sun, Y.; Qin, J.; Yang, X.; Zhang, P.; Wang, Q.; Zou, L.; Zhang, Y.; Zhang, L.; Fang, Y.; Li, J.; Zhou, H. C. Stable metal−organic frameworks: design, synthesis, and applications. Adv. Mater. 2018, 30, 1704303. (37) Huang, H. L.; Li, J. R.; Wang, K. K.; Han, T. T.; Tong, M. M.; Li, L. S.; Xie, Y. B.; Yang, Q. Y.; Liu, D. H.; Zhong, C. L. An in situ self-assembly template strategy for the preparation of hierarchicalpore metal-organic frameworks. Nat. Commun. 2015, 6, 8847. (38) Li, H.; Eddaoudi, M.; O’Keeffe, M.; Yaghi, O. M. Design and synthesis of an exceptionally stable and highly porous metal-organic framework. Nature 1999, 402 (6759), 276−279. (39) Yuan, S. A.; Zou, L. F.; Qin, J. S.; Li, J. L.; Huang, L.; Feng, L. A.; Wang, X. A.; Bosch, M.; Alsalme, A.; Cagin, T.; Zhou, H. C. Construction of hierarchically porous metal-organic frameworks through linker labilization. Nat. Commun. 2017, 8, 15356. (40) Sun, L. B.; Li, J. R.; Park, J.; Zhou, H. C. Cooperative templatedirected assembly of mesoporous metal-organic frameworks. J. Am. Chem. Soc. 2012, 134 (1), 126−129. (41) Boissonnault, J. A.; Wong-Foy, A. G.; Matzger, A. J. Core-shell structures arise naturally during ligand exchange in metal-organic frameworks. J. Am. Chem. Soc. 2017, 139 (42), 14841−14844. (42) Chen, Y. Z.; Wang, Z. U.; Wang, H. W.; Lu, J. L.; Yu, S. H.; Jiang, H. L. Singlet oxygen-engaged selective photo-oxidation over Pt nanocrystals/porphyrinic MOF: the roles of photothermal effect and Pt electronic state. J. Am. Chem. Soc. 2017, 139 (5), 2035−2044.

H

DOI: 10.1021/acscentsci.8b00722 ACS Cent. Sci. XXXX, XXX, XXX−XXX