Controlling the Orientation of Metal–Organic Framework Crystals by

Dec 5, 2017 - Synopsis. Highly oriented three-dimensional metal−organic framework (MOF) crystals with a pillared-layer structure were prepared on a ...
0 downloads 14 Views 3MB Size
Subscriber access provided by READING UNIV

Article

Controlling the Orientation of Metal-Organic Framework Crystals by an Interfacial Growth Approach Using a Metal Ion-Doped Polymer Substrate Takashi Ohhashi, Takaaki Tsuruoka, Seiya Fujimoto, Yohei Takashima, and Kensuke Akamatsu Cryst. Growth Des., Just Accepted Manuscript • DOI: 10.1021/acs.cgd.7b01402 • Publication Date (Web): 05 Dec 2017 Downloaded from http://pubs.acs.org on December 5, 2017

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

Crystal Growth & Design 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 25 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

Crystal Growth & Design

Controlling the Orientation of Metal-Organic Framework Crystals by an Interfacial Growth Approach Using a Metal Ion-Doped Polymer Substrate Takashi Ohhashi, Takaaki Tsuruoka*, Seiya Fujimoto, Yohei Takashima, and Kensuke Akamatsu* Department of Nanobiochemistry, Frontiers of Innovative Research in Science and Technology (FIRST), Konan University, 7-1-20 Minatojimaminami, Chuo-ku, Kobe 650-0047, Japan KEYWORDS Metal-organic frameworks, Orientation control, Interfacial reaction

ABSTRACT (Word Style “BD_Abstract”).

Highly oriented three-dimensional metal-organic framework (MOF) crystals with a pillared-layer structure were prepared on a metal ion-doped polymer substrate. This approach allowed the formation of MOF crystals with controlled crystalline orientation in a one-pot reaction facilitated by the control of the chemical interaction between the framework components and polymer

ACS Paragon Plus Environment

1

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

Page 2 of 25

substrate. Further, this approach provides a new concept for the one-pot synthesis of oriented MOF crystals and furthers the fundamental understanding of the crystal growth mechanism based on self-assembly processes on two-dimensional support substrates.

TEXT. Introduction Metal-organic frameworks (MOFs) have attracted considerable attention as a class of highly ordered porous crystal materials with unique properties and potential applications in gas storage1,2, separation3,4, catalysis5,6, and sensing7,8. Although studies on MOFs have been mainly focused on bulk materials, recently, there have been increasing efforts on preparing MOF crystals on substrates for controlling the pore space and integrating specific functions9-13. In terms of their fabrication, the ability to control the crystalline orientation on a substrate is very important to realize desired properties for more advanced applications such as selective gas separation and chemical sensing14-16. However, it is difficult to fabricate oriented MOF crystals on substrates because the spontaneous crystal growth process from a solution containing MOF components leads to the formation of randomly oriented MOF crystals17,18. To induce oriented self-assembly as well as the formation of MOF crystal-based multi-layers, a layer-by-layer (LbL) synthesis approach using self-assembled monolayer (SAM)-functionalized substrates has been intensively studied19-22. Despite its effectiveness and versatility, the method relies on multi-step processes involving the formation of SAM and LbL growth of MOF crystals through metal adsorption/ligand adsorption steps. To date, although there have been few reports on successful orientation control of MOF crystals on SAM-modified substrates in a one-pot reaction system23,

ACS Paragon Plus Environment

2

Page 3 of 25 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

Crystal Growth & Design

this approach presents significant limitations in terms of fabricating uniform and continuous MOF crystal-based films because the reaction solution is too dilute to produce multi-layers. Therefore, it is important to develop a straightforward methodology for constructing highly oriented MOF crystal-based multi-layers on a substrate with control over the self-assembly process in a one-pot synthesis, both to further the fundamental understanding of the crystal growth mechanism based on self-assembly processes and to realize more advanced applications based on MOF crystals. Herein, we demonstrate a new concept for a one-pot synthesis of highly oriented MOF crystals based on the control of chemical interactions between the framework components and support substrate. As a proof-of-concept, we focus on an anisotropic tetragonal framework, [Cu2(ndc)2dabco]n

(ndc

=

1,4-naphthalene

dicarboxylate;

dabco

=

1,4-

diazabicyclo[2.2.2]octane), to fabricate MOF crystals on a substrate, because it has a simple three-dimensional (3-D) porous network with a pillared-layer structure, where the 2-D layered grids consisting of Cu-ndc paddle-wheel units are connected by dabco pillar ligands. The tetragonal framework exhibits a rectangular prism morphology in which the four surfaces endcapped by the layered carboxylic acid groups are denoted as (100) surfaces, and the other two surfaces terminated by nitrogen pillar ligands are denoted as (001) surfaces24-27. The crystalline orientation of [Cu2(ndc)2dabco]n on the substrate are expected to depend on the direction of the 2-D layer with respect to the substrate in the initial reaction stage because the construction of the framework is based on the connection of the layers by the pillar ligands. We have recently succeeded in controlling the crystal growth of the [Cu2(ndc)2dabco]n framework using an interfacial growth approach with a metal ion-doped polymer substrate28. In this approach, the ion-exchange reaction between the doped metal ions in the polymer substrate and organic ligands

ACS Paragon Plus Environment

3

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

Page 4 of 25

bearing carboxylic acid groups acts as a trigger for the crystal growth, which allows the selfassembly process of [Cu2(ndc)2dabco]n crystals to be separated from the formation of Cu-ndc paddle-wheel unit-based layers and linkage of the pillar ligands in a one-pot reaction system containing both H2bdc and dabco ligands. The sequentially controlled MOF crystal growth process would allow control over the direction of the 2-D layers on the substrate by modulating the synthetic parameters, in which the 2-D layers parallel or vertical with respect to the substrate could provide [001] or [100] orientation, respectively (Scheme 1). In this work, we attempted to control the direction of the 2-D layers by controlling the chemical interaction between the 2-D layers and polymer substrate by modulating the layer size (contact area with the substrate) and solvent environment, to achieve control over the orientation of the [Cu2(ndc)2dabco]n crystals.

Experimental Section Potassium hydroxide, copper(II) acetate monohydrate, H2ndc, and dabco were purchased from Wako Chemicals Ltd. Methanol, ethanol, and n-butanol were purchased from Kanto Chemical Ltd. Pyromellitic dianhydride oxydianiline (PMDA-ODA) type polyimide films (50 µm thick, Kapton 200H, Toray-Du Pont Co. Ltd.) were used as the polymer substrate. The films were cleaned with ethanol before use. The polyimide films (1 × 2 cm2) were initially immersed in a 5 M KOH solution at 50 °C for 5 min, followed by through rinsing with copious amounts of distilled water. The modified films were then immersed in a 100 mM aqueous Cu(CH3COO)2 solution at room temperature for 20 min. After rinsing with distilled water, the Cu2+-doped polymer films were immersed in methanol or n-butanol solution (5 mL) consisting of H2ndc and dabco (10 and 5 mM,

ACS Paragon Plus Environment

4

Page 5 of 25 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

Crystal Growth & Design

respectively), followed by heating at 150 °C for 30 min by microwave irradiation (Initiator+; Biotage). The crystal morphology and size of the obtained samples were observed by scanning electron microscopy (SEM; JSM-7001FA, JEOL). The crystal orientation was evaluated by X-ray diffraction (XRD) measurements (RINT-2200 Ultima IV, Rigaku) using Cu Kα radiation.

Results and Discussion As described in the experimental section, [Cu2(ndc)2dabco]n crystals were prepared on the Cu2+ ion-adsorbed polyimide film immersed in methanol solution, using microwave irradiation. SEM observation of the obtained samples indicated the deposition of adequately intergrown polycrystalline MOFs with tetragonal morphology on the substrate (Figure 1a). The XRD measurement of the resulting product indicated the tetragonal crystallization with the diffraction pattern being identical to the simulated pattern of the [Cu2(ndc)2dabco]n crystal and the halo pattern of polyimide film (Figure 1b). Although the crystal morphology of the obtained sample is tetragonal, the relative intensity of the two characteristic peaks corresponding to the (001) and (100) planes, I(001)/I(100), is evidently larger in comparison with that of the simulated pattern. Thus, the obtained [Cu2(ndc)2dabco]n crystals are oriented along the [001] direction on the substrate, which suggests that the 2-D layers are aligned parallel to the polymer substrate, in the initial reaction stage. To elucidate the factors that induce the formation of [Cu2(ndc)2dabco]n crystals oriented along the [001] direction on the polymer substrate in this synthesis system, we investigated the effect

ACS Paragon Plus Environment

5

Crystal Growth & Design 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 6 of 25

of the ligand concentration. First, H2ndc ligand concentration was changed in the range of 1–40 mM, while that of the dabco ligand was maintained constant at 5 mM. As shown in Figure 2a, the relative peak intensity, I(001)/I(100), gradually increased with increasing H2ndc ligand concentration. Considering that the tetragonal morphology remains unchanged in all the samples (Figure 2b), the increase in the relative peak intensity implies that the [Cu2(ndc)2dabco]n crystals are better oriented along the [001] direction at higher H2ndc ligand concentration. Additionally, in our previous study, we revealed that increasing the H2ndc concentration leads to an increase in the size of the 2-D layer, because the elution of the doped metal ions from the substrate is based on an ion-exchange reaction between the metal ions and protons of the H2ndc molecules27, that is, the [Cu2(ndc)2dabco]n crystals are highly oriented along the [001] direction when the size of the 2-D layer is large. Next, we conducted the synthesis with varying dabco ligand concentration, while the H2ndc ligand concentration was maintained at 40 mM, the condition that induces the growth of highly oriented crystals along the [001] direction. Although the size and morphology are almost same under all reaction conditions, the relative peak intensity, I(001)/I(100), obviously increased with increasing dabco ligand concentration in the range of 2.5–30 mM (Figure 3), which indicates that increasing the dabco concentration facilitates the growth of highly oriented [Cu2(ndc)2dabco]n crystals along the [001] direction. As an excess amount of dabco ligands can facilitate faster connection between the 2-D layers at the lattice points, the layers could maintain a parallel orientation on the polymer substrate, leading to the growth of [Cu2(ndc)2dabco]n crystals oriented along the [001] direction. On the contrary, the crystalline orientation was hardly dependent on the dabco ligand concentration when H2ndc ligand concentration was maintained at 5 mM (Figure S1). These results suggest that H2ndc ligand concentration is a dominant factor for

ACS Paragon Plus Environment

6

Page 7 of 25 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

Crystal Growth & Design

the growth of [Cu2(ndc)2dabco]n crystals oriented along the [001] direction, whereas the dabco ligand concentration influences the crystal orientation at only high H2ndc concentration. To gain more insight to the crystalline orientation of the [Cu2(ndc)2dabco]n crystals, we evaluated the effect of the solvent, because this significantly influences the chemical interactions between the 2-D layer units and the polyimide film, which might direct the crystal growth of the 2-D layers along a different direction with respect to the substrate. Hence, the [Cu2(ndc)2dabco]n crystals were prepared in n-butanol solution which is an alcoholic solvent similar to methanol, however, with a polarity lower than that of methanol. When the synthesis was conducted under n-butanol solution consisting 40 mM H2ndc and 30 mM dabco ligands, conditions that lead to [Cu2(ndc)2dabco]n crystals highly oriented along the [001] direction in methanol solution, the relative peak intensity, I(001)/I(100), dropped below 1 (Figure 4a), implying that the [Cu2(ndc)2dabco]n crystals prepared in butanol solution are oriented along the [100] direction in contrast with the orientation realized in a methanolic solution. Moreover, the crystal morphology of the sample obtained in butanol solution is plate-like (Figure 4b), in which the crystal growth along the [100] direction is preferable. Therefore, the different relative peak intensity observed is likely due to not only the crystalline orientation along the [100] direction but also the crystal morphology. To elucidate the factors that influence the relative peak intensity in the butanol system, we investigated the effect of the ligand concentration. When the [Cu2(ndc)2dabco]n crystals were prepared with varying H2ndc ligand concentration while maintaining the dabco ligand concentration at 5 mM, the relative peak intensity, I(100)/I(001), increased with H2ndc ligand concentration in the range of 1–20 mM (Figure 5a). The obtained SEM images show that the plate-like crystal morphology remains unchanged, while the crystal size increase with increasing

ACS Paragon Plus Environment

7

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

Page 8 of 25

H2ndc concentration (Figure 5b). These results imply that the variation in the relative peak intensity with varying H2ndc ligand concentration is most likely attributed to the formation of crystals highly oriented along the [100] direction as distinct from the influence of crystal morphology, because the isotropic increase of the crystal size essentially lead to the increase of the only peak intensity while the relative peak intensity keeps constant. From the SEM images, it is also obvious that the deposited [Cu2(ndc)2dabco]n crystals are sparse on the substrate for low concentration range of the H2ndc ligand. This is due to the slow elution rate of the doped-metal ions from the substrate, because of the low solubility of the metal ions in butanol solution; this induces slower ion-exchange rate at lower H2ndc concentration as the ion-exchange occurs between the doped-metal ions and the protons of the H2ndc molecules. These results suggest that the 2-D layers could align vertically on the substrate in the butanol system when the nuclei are formed densely in the initial reaction stage, leading to the growth of highly oriented [Cu2(ndc)2dabco]n crystals along the [100] direction at high H2ndc ligand concentrations. Next, we conducted the synthesis with varying dabco ligand concentration, while maintaining the concentration of the H2ndc ligand at 20 mM, that is a suitable condition to grow highly oriented crystals along the [100] direction. Although the size and morphology of the crystals are almost same under all the reaction conditions, the relative peak intensity, I(100)/I(001), decreased with increasing dabco ligand concentration (Figure 6), indicating that a low dabco concentration leads to the growth of [Cu2(ndc)2dabco]n crystals highly oriented along the [100] direction. In these cases, faster connection of the 2-D layers by an excess amount of the pillar ligands could induce the crystal growth in the sparse state of the 2-D layers due to the low elution rate of the dopedmetal ions in butanol solution, thereby leading to the growth of randomly oriented [Cu2(ndc)2dabco]n crystals. Therefore, the rapid connection of layers by dabco ligands might be

ACS Paragon Plus Environment

8

Page 9 of 25 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

Crystal Growth & Design

unfavorable for the preparation of [Cu2(ndc)2dabco]n crystals oriented along the [100] direction, because the 2-D layers could be aligned vertically with respect to the substrate when the nuclei are formed densely in the initial reaction stage. On the contrary, the crystalline orientation was hardly dependent on the dabco ligand concentration when the H2ndc ligand concentration was maintained at 5 mM (Figure S2). These results correspond with the trend in the methanol system, that is, H2ndc ligand concentration is a dominant factor for the growth of highly oriented [Cu2(ndc)2dabco]n crystals, whereas the dabco ligand concentration influences the orientation at only high H2ndc concentration. The effect of solvent on the crystalline orientation of [Cu2(ndc)2dabco]n frameworks was investigated further using a mixed solvent consisting of methanol and butanol. As shown Figure S3a, the relative peak intensity varied with the ratio of methanol and butanol. The relative peak intensity, I(001)/I(100), gradually decreased with decreasing methanol content, and finally, the crystals obtained under higher butanol content showed [100] orientation on the substrate. This variation in the relative peak intensity is likely due to the polarity of the reaction solution, with the higher and lower polarity alcohol solvent facilitating the growth of the [Cu2(ndc)2dabco]n crystals oriented along the [001] and [100] directions, respectively. In addition, the resulting SEM images of MOFs grown in different solvent compositional ratio indicate that the crystal size decreases with decreasing methanol content (Figure S3b). This is due to decreased elution rate of the metal ions doped in the polyimide film with increasing butanol content. We have also prepared [Cu2(ndc)2dabco]n crystals in ethanol and propanol solution in order to investigate the effect of polarity of the alcohol solvent for orientation control in the present approach, in which the concentration of H2ndc and dabco is 20 mM and 5 mM, respectively. The relative peak intensity, I(001)/I(100), drastically decreased by change in solvent from methanol to ethanol, and

ACS Paragon Plus Environment

9

Crystal Growth & Design 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 25

then, gradually decreased with increasing alkyl chain (Figure S4a). This variation is caused by the higher polarity of methanol in comparison to that of ethanol, propanol, and butanol. Therefore, the MOF crystals are oriented along [001] direction in methanol solution with higher polarity, while the solvents with the lower polarity leads to the [100] orientation of the obtained crystals on the substrate. Additionally, the crystal size of the obtained samples decreased with decreasing polarity of alcohol solvent because of slower elution rate of doped metal ions (Figure S4b). These results well correspond to those of the reaction conditions using a mixed solvent consisting of methanol and butanol. The orientation of the [Cu2(ndc)2dabco]n crystals on the polymer substrate could be easily controlled by modulating the polarity of the alcohol solvent in the proposed approach. Herein, to confirm whether the orientation of the [Cu2(ndc)2dabco]n crystals may be determined based on the direction of the 2-D layers with respect to the substrate in the initial reaction stage, a time course analysis of the crystal growth was performed through SEM observations and XRD measurements. In the case of butanol system, a certain amount of the crystals formed on the polymer substrate in the initial stage of the reaction during 5 min at an H2ndc ligand concentration of 20 mM, with the relative peak intensity, I(100)/I(001), being 7.4 (Figure S5) indicating that the obtained [Cu2(ndc)2dabco]n crystals are oriented along the [100] direction. During the reaction for 10 min, the crystals are more densely deposited, causing a slight increase in the relative peak intensity in the range of 5 and 10 min. After a 10 min reaction time, as only the crystal size increased isotropically, the relative peak intensity is almost the same for samples prepared during this time. These results imply that the 2-D layers aligned vertically on the substrate, followed by the growth of the crystals with vertically oriented 2-D layers. In contrast, the crystals are densely deposited in the initial stage of the reaction during 5

ACS Paragon Plus Environment

10

Page 11 of 25 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

Crystal Growth & Design

min in the case of methanol system (Figure S6). This is attributed to a sufficiently fast elution rate of the doped-metal ions, which leads to fast nucleation and crystal growth. SEM observations indicate that only the crystal size increased isotropically with increasing reaction time. Therefore, the relative peak intensity, I(001)/I(100), remained constant for the crystals obtained during 5 to 30 min (Figure S6). These results imply that in the methanol solution, crystals with the 2-D layers parallel to the substrate are densely deposited in the reaction stage during 5 min, followed by crystal growth. These compelling results imply that the 2-D layers prefer to lie flat on the polymer substrate in methanol solution and to align vertically on the substrate in butanol solution. This difference may be caused by the chemical interactions between the 2-D layer units and polyimide film, which is speculated to mainly involve π-π stacking interactions because the layers consist of naphthalene rings and the polyimide substrate contains numerous aromatic rings. As the stacking interaction is more effective in methanol solution owing to its high polarity, the 2-D layers are prone to align parallel to the polymer substrate, resulting in the formation of [Cu2(ndc)2dabco]n crystals oriented along the [001] direction. On the contrary, the stacking effect is weak in butanol solution owing to its low polarity, which leads to random or vertical orientation of the 2-D layers on the substrate. On the basis of this consideration and the obtained results, we propose a mechanism for the orientation control of the [Cu2(ndc)2dabco]n crystals upon varying the solvent and H2ndc ligand concentration (Scheme 2). When methanol is used as the solvent, the 2-D layers formed in the initial reaction stage prefer to lie flat on the substrate due to the strong interaction with the polyimide film, which leads to the formation of [Cu2(ndc)2dabco]n crystals oriented along the [001] direction. In addition, large-sized 2-D layers form when the concentration of the H2ndc ligands is high, because the elution rate of the doped-metal ions is

ACS Paragon Plus Environment

11

Crystal Growth & Design 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 12 of 25

faster. Consequently, the larger 2-D layers prefer lie parallel to the substrate because of the increased contact area with the polymer film, resulting in the growth of highly oriented crystals along the [001] direction. On the contrary, the 2-D layers tend to align vertically on the substrate in butanol system because the chemical interaction between the layer units and polyimide film is weak, which leads to the growth of [Cu2(ndc)2dabco]n crystals oriented along the [100] direction. Additionally, the [Cu2(ndc)2dabco]n crystals are highly oriented along the [100] direction because of dense nucleation when the H2ndc ligand concentration is high. Overall, the direction of the 2-D layers could be controlled by modulating the chemical interaction between the layer units and polymer film units in the present approach, with the type of solvent and concentration of the H2ndc ligands being the key parameters. The present approach was also applied to the orientation control of other anisotropic tetragonal framework, [Cu2(bdc)2dabco]n (bdc = 1,4-benzenedicarboxylate) crystal. This framework shows same pillared-layer structure as [Cu2(ndc)2dabco]n crystals, where the 2-D layered grids consisting of Cu-bdc paddle-wheel units are connected by dabco pillar ligands. The XRD measurement reveals that the [Cu2(bdc)2dabco]n crystals obtained in methanol solution show high relative peak intensity, I(001)/I(100), of 2.54, while the ratio drop below 1.0 in butanol system (Figure S7a). Considering that the crystal morphology of both samples obtained methanol and butanol solution is tetragonal from SEM images (Figure S7b), it is found that the obtained [Cu2(bdc)2dabco]n crystals in methanol are oriented along [001] and the samples obtained in butanol solution show random orientation on the polymer substrate. These results demonstrated that the present approach for orientation control based on the control of the chemical interaction between framework components and polymer substrate can be readily applied to other pillaredlayer MOF crystals.

ACS Paragon Plus Environment

12

Page 13 of 25 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

Crystal Growth & Design

Conclusions We demonstrated a one-pot synthesis of highly oriented [Cu2(ndc)2dabco]n crystals via an interfacial growth approach using a metal ion-doped polymer substrate. Modulation of the chemical interactions between the framework components and polymer units allowed control over the orientation of the 2-D Cu-ndc layers formed on the substrate, leading to the growth of [Cu2(ndc)2dabco]n crystals oriented along [001] or [100] direction. Moreover, the controllable ion-exchange reaction rate facilitated by the H2ndc ligand concentration plays a crucial role in controlling the crystal orientation. The successful approach based on fundamental science such as the kinetic control of the doped-metal ions and modulation of the chemical interaction between the 2-D framework units and polymer substrate would pave the way for one pot syntheses of highly oriented MOF crystals in the future.

ACS Paragon Plus Environment

13

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

Page 14 of 25

FIGURES

Figure 1. (a) SEM image and (b) experimental and simulated XRD patterns of [Cu2(ndc)2dabco]n crystals and polyimide film.

Figure 2. (a) XRD patterns and relative peak intensities, I(001)/I(100), with respect to H2ndc concentration, and (b) SEM images of [Cu2(ndc)2dabco]n crystals obtained at H2ndc ligand concentrations of 1–40 mM. The concentration of the dabco ligand was fixed at 5 mM.

ACS Paragon Plus Environment

14

Page 15 of 25 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

Crystal Growth & Design

Figure 3. (a) XRD patterns and relative peak intensities, I(001)/I(100), with respect to the concentration of dabco, and (b) SEM images of [Cu2(ndc)2dabco]n crystals obtained at dabco ligand concentrations of 2.5–30 mM. The concentration of H2ndc ligand was fixed at 40 mM.

Figure 4. (a) XRD patterns and (b) SEM image for the [Cu2(ndc)2dabco]n crystals prepared in butanol solution at 40 mM H2ndc and 30 mM dabco ligand.

ACS Paragon Plus Environment

15

Crystal Growth & Design 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 16 of 25

Figure 5. (a) XRD patterns and relative peak intensities, I(100)/I(001), with respect to H2ndc concentration, and (b) SEM images of [Cu2(ndc)2dabco]n crystals obtained at H2ndc ligand concentrations of 1–20 mM. The concentration of dabco ligand was fixed at 5 mM.

Figure 6. (a) XRD patterns and relative peak intensities, I(100)/I(001), with respect to dabco concentration and (b) SEM images of [Cu2(ndc)2dabco]n crystals obtained at dabco ligand concentrations of 2.5–30 mM. The concentration of H2ndc ligand was fixed at 20 mM.

ACS Paragon Plus Environment

16

Page 17 of 25 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

Crystal Growth & Design

SCHEMES. Scheme 1. Schematic illustration of the strategy used for controlling the orientation of [Cu2(ndc)2dabco]n crystals on a 2-D substrate.

Scheme 2. Schematic representation of the proposed strategy for the growth of highly oriented [Cu2(ndc)2dabco]n crystals in a one-pot reaction by controlling of the chemical interactions between the framework components and polyimide film.

ACS Paragon Plus Environment

17

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

Page 18 of 25

ASSOCIATED CONTENT Supporting Information. SEM images and XRD patterns of the obtained MOFs by changing the concentration of dabco ligands in methanol system; SEM images and XRD patterns of the obtained MOFs by changing the concentration of dabco ligands in butanol system; SEM images and XRD patterns of the obtained MOFs by using a mixed solvent consisting of methanol and butanol; Analysis of time course of reaction in butanol solution by SEM observation and XRD measurement; Analysis of time course of reaction in methanol solution by SEM observation and XRD measurement.

AUTHOR INFORMATION Corresponding Author *(T.T.) E-mail: [email protected] *(K.A.) E-mail: [email protected]

ACKNOWLEDGMENT This work was supported by JSPS KAKENHI Grant Number 15K05655. This work was also supported by “Molecular Technology” of Strategic International Collaborative Research Program (SICORP) from the Japan Science and Technology Agency (JST)-Agence Nationale de la Recherche (ANR).

ACS Paragon Plus Environment

18

Page 19 of 25 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

Crystal Growth & Design

REFERENCES 1) Kondo, M.; Yoshitomi, T.; Seki, K.; Matsuzaka, H.; Kitagawa, S. Three-Dimensional Framework with Channeling Cavities for Small Molecules: {[M2(4, 4′-bpy)3(NO3)4]·xH2O}n (M = Co, Ni, Zn). Angew. Chem., Int. Ed. 1997, 36, 1725–1727.

2) Rowsell, J. L. C.; Millward, A. R.; Park, K. S.; Yaghi, O. M. Hydrogen Sorption in Functionalized Metal-Organic Frameworks. J. Am. Chem. Soc. 2004, 126, 5666–5667.

3) Mueller, U.; Schubert, M.; Teich, F.; Puetter, H.; Schierle-Arndt, K.; Pastre, J. Metal-organic frameworks-prospective industrial applications. J. Mater. Chem. 2006, 16, 626–636.

4) Couck, S.; Denayer, J. F. M.; Baron, G. V.; Remy, T.; Gascon, J.; Kapteijn, F. An AmineFunctionalized MIL-53 Metal-Organic Framework with Large Separation Power for CO2 and CH4. J. Am. Chem. Soc. 2009, 131, 6326–6327.

5) Horike, S.; Dincá, M. Tamaki, K.; Long, J. R. Size-Selective Lewis Acid Catalysis in a Microporous Metal-Organic Framework with Exposed Mn2+ Coordination Sites. J. Am. Chem. Soc. 2008, 130, 5854–5855.

ACS Paragon Plus Environment

19

Crystal Growth & Design 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 20 of 25

6) Lee, J.; Farha, O. K.; Roberts, J.; Scheidt. K. A.; Nguyen. S. T.; Hup, J. T. Metal-organic framework materials as catalysts. Chem. Soc. Rev. 2009, 38, 1450–1459.

7) Chen, B.; Wang, L.; Xiao, Y.; Fronczek, F. R.; Xue, M.; Cui, Y.; Qian, G. A luminescent metal-organic framework with Lewis basic pyridyl sites for the sensing of metal ions. Angew. Chem., Int. Ed. 2009, 48, 500–503.

8) Kreno, L. E.; Leong, K.; Farha, O. K.; Allendorf, M.; Van Duyne, R. P.; Hupp, J. T. Metalorganic framework materials as chemical sensors. Chem. Rev. 2012, 112, 1105–1125.

9) Pan, Y.; Lai, Z. Sharp Separation of C2/C3 Hydrocarbon Mixtures by Zeolitic Imidazolate Framework-8 (ZIF-8) Membranes Synthesized in Aqueous Solutions. Chem. Commun. 2011, 47, 10275–10277.

10) Aguado, S.; Nicolas, C.-H.; Moizan-Baslé, V.; Nieto, C.; Amrouche, H.; Bats, N; Audebrand, N.; Farrussseng, D. Facile synthesis of an Ultramicroporous MOF Tubular Membrane with Selectivity Towards CO2. New J. Chem. 2011, 35, 41–44.

ACS Paragon Plus Environment

20

Page 21 of 25 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

Crystal Growth & Design

11) Shah, M. N.; Gonzales, M. A.; McCarthy, M. C.; jeong, H.-K. An Unconventional Rapid Synthesis of High Performance Metal-Organic Framework Membranes. Langmuir 2013, 29, 7896–7902.

12) Meilikhov, M.; Furukawa, S.; Hirai, K.; Fischer, R. A.; Kitagawa, S. Binary Janus Porous Coordination Polymer Coatings for Sensor Devices with tunable Analyte Affinity. Angew. Chem., Int. Ed. 2013, 52, 341–345.

13) Sakaide, S.; Otsubo, K.; Sakata, O.; Song, C.; Fujiwara, A.; Takata, M.; Kitagawa, H.; Crystalline Coordination Framework Endowed with Dynamic Gate-Opening Behaviour by Being Down-Sized to a Thin Film. Nat. Chem. 2016, 8, 377–383.

14) Bux, H.; Feldhoff, A.; Cravillon, J.; Wiebcke, M.; Li, Y.-S.; Caro. J. Oriented Zeolitic Imidazolate Framework-8 (ZIF-8) Membrane with Sharp H2/C3H8 Molecular Sieve Separation. Chem. Mater. 2011, 23, 2262–2269.

15) Falcaro, P.; Okada, K.; Hara, T.; Ikigaki, K.; Tokudome, Y.; Thornton, A. W.; Hill, A. J.; Williams, T.; Doonan, C.; Takahashi, M. Centimetre-scale micropore alignment in oriented polycrystalline metal-organic framework films via heteroepitaxial growth. Nat. Mater. 2017, 16, 342–348.

ACS Paragon Plus Environment

21

Crystal Growth & Design 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 22 of 25

16) Liu, Y.; Zeng, G.; Pan, Y.; Lai, Z. Synthesis of highly c-oriented ZIF-69 membranes by secondary growth and their gas permeation properties. J. Membrane Sci. 2011, 379, 46-51

17) Hermes, S.; Schröder, F.; Chelmowski, R.; Wöll, C.; Fischer, R. A. Selective Nucleation and Growth of Metal-Organic Open framework Thin Films on Patterned COOH/CF3-Terminated Self-Assembled Monolayers on Au(111). J. Am. Chem. Soc. 2005, 127, 13744–13745.

18) Zacher, D.; Baunemann, A.; Hermes, S.; Fischer, R. A. Deposition of microcrystalline [Cu3(btc)2] and [Zn2(bdc)2(dabco)] at alumina and silica surfaces modified with patterned self assembled organic monolayers: evidence of surface selective and oriented growth. J. Mater. Chem. 2007, 17, 2785–2792.

19) Shekhah, O.; Wang, H.; Zacher, D.; Fischer, R. A.; Wöll, C. Growth Mechanism of MetalOrganic Frameworks: Insights into the Nucleation by Employing a Step-by-Step Route. Angew. Chem., Int. Ed. 2009, 48, 5038–5041.

20) Liu, B.; Shekhah, O.; Arslan, H. K.; Liu, J.; Wöll, C.; Fischer, R. A. Enantiopure MetalOrganic Framework Thin Films: Oriented SURMOF Growth and Enantioselective Adsorption. Angew. Chem., Int. Ed. 2012, 51, 807–810.

ACS Paragon Plus Environment

22

Page 23 of 25 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

Crystal Growth & Design

21) Otubo, K.; Haraguchi, T.; Sakata, O.; Fujiwara, A.; Kitagawa, H. Step-by-Step Fabrication of a Highly Oriented Crystalline Three-Dimensional Pillared-Layer-Type Metal-Organic Framework Thin Film Confirmed by Synchrotron X-ray Diffraction. J. Am. Chem. Soc. 2012, 134, 9605–9608.

22) Chen, S.-M.; Liu, M.; Gu, Z.-G.; Fu, W.-Q.; Zhang, J. Chiral Chemistry of Homochiral Porous Thin Film with Different Growth Orientations. ACS Appl. Mater. Interfaces 2016, 8, 27332–27338.

23) Biemmi, E.; Scherb, C.; Bein, T. Oriented Growth of the Metal Organic Framework Cu3(BTC)2(H2O)3·xH2O Tunable with Functionalized Self-Assembled Monolayers. J. Am. Chem. Soc. 2007, 129, 8054–8055.

24) Dybtsev, D. N.; Chun, H.; Kim, K. Rigid and Flexible: A Highly Porous Metal-Organic Framework with Unusual Guest-Dependent Dynamic Behavior. Angew. Chem., Int. Ed. 2004, 43, 5033–5036.

ACS Paragon Plus Environment

23

Crystal Growth & Design 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 24 of 25

25) Kitaura, R.; Iwahori, F.; Matsuda, R.; Kitagawa, S.; Kubota, Y.; Takata, M.; Kobayashi, T. C. Rational Design and Crystal Structure Determination of a 3-D Metal-Organic Jungle Gymlike Open Framework. Inorg. Chem. 2004, 43, 6522–6524.

26) 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, 1766–1770.

27) Kondo, M.; Furukawa, S.; Hirai, K.; Kitagawa, S. Coordinatively Immobilized Monolayers on Porous Coordination Polymer Crystals. Angew. Chem., Int. Ed. 2010, 49, 5327–5330.

28) Tsuruoka, T.; Mantani, K.; Miyanaga, A.; Matsuyama, T.; Ohhashi, T.; Takashima, Y.; Akamatsu, K. Morphology Control of Metal-Organic Frameworks Based on Paddle-Wheel Units on Ion-Doped Polymer Substrate Using an Interfacial Growth Approach. Langmuir 2016, 32, 6068–6073.

For Table of Contents Use only

ACS Paragon Plus Environment

24

Page 25 of 25 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

Crystal Growth & Design

Manuscript Title: Controlling the Orientation of Metal-Organic Framework Crystals by an Interfacial Growth Approach Using a Metal Ion-Doped Polymer Substrate

Author List: Takashi Ohhashi, Takaaki Tsuruoka,* Seiya Fujimoto, Yohei Takashima, and Kensuke Akamatsu* Department of Nanobiochemistry, Frontiers of Innovative Research in Science and Technology (FIRST), Konan University, 7-1-20 Minatojimaminami, Chuo-ku, Kobe 650-0047, Japan

TOC Graphic

Synopsis Highly oriented three-dimensional metal-organic framework (MOF) crystals with a pillared-layer structure were prepared on a metal ion-doped polymer substrate. This approach allowed the formation of MOF crystals with controlled crystalline orientation in a one-pot reaction facilitated by the control of the chemical interaction between the framework components and polymer substrate.

ACS Paragon Plus Environment

25