Multifaceted Modularity: A Key for Stepwise Building of Hierarchical


Multifaceted Modularity: A Key for Stepwise Building of Hierarchical...

0 downloads 72 Views 4MB Size

Subscriber access provided by University of Florida | Smathers Libraries

Article

Multifaceted Modularity: a Key for Stepwise Building of Hierarchical Complexity in An-MOFs Ekaterina A. Dolgopolova, Otega A. Ejegbavwo, Corey R. Martin, Mark D. Smith, Wahyu Setyawan, Stavros Karakalos, Charles Henry Henager, Hans-Conrad zur Loye, and Natalia B. Shustova J. Am. Chem. Soc., Just Accepted Manuscript • DOI: 10.1021/jacs.7b09496 • Publication Date (Web): 25 Oct 2017 Downloaded from http://pubs.acs.org on October 25, 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.

Journal of the American Chemical Society 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 21

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

Journal of the American Chemical Society

Multifaceted Modularity: a Key for Stepwise Building of Hierarchical Complexity in An-MOFs Ekaterina A. Dolgopolova,a Otega A. Ejegbavwo,a Corey R. Martin,a Mark D. Smith,a Wahyu Setyawan,c Stavros Karakalos,b Charles Henry Henager,c Hans-Conrad zur Loye,a and Natalia B. Shustovaa* a

Department of Chemistry and Biochemistry, University of South Carolina, SC 29208, United States College of Engineering and Computing, University of South Carolina, SC 29208, United States c Pacific Northwest National Laboratory, Richland, WA 99352, United States b

ABSTRACT: Growing necessity for efficient nuclear waste management is a driving force for development of alternative architectures towards fundamental understanding of mechanisms involved in actinide integration inside extended structures. In this manuscript, metal-organic frameworks (MOFs) were investigated as a model system for engineering radionuclide containing materials through utilization of unprecedented MOF modularity, which cannot be replicated in any other type of materials. Through the implementation of recent synthetic advances in the MOF field, hierarchical complexity of An-materials was built stepwise, which was only feasible due to preparation of the first examples of actinide-based frameworks with “unsaturated” metal nodes. The first successful attempts of solid-state metathesis and metal node extension in An-MOFs are reported, and the results of the former approach revealed drastic differences in chemical behavior of extended structures versus molecular species. Successful utilization of MOF modularity also allowed us to structurally characterize the first example of bimetallic An-An nodes. To the best of our knowledge, through combination of solid-state metathesis, guest incorporation, and capping linker installation, we were able to achieve the highest Th wt% in mono- and bi-actinide frameworks with minimal structural density. Overall, the combination of a multistep synthetic approach with homogeneous actinide distribution and moderate solvothermal conditions could make MOFs an exceptionally powerful tool to address fundamental questions responsible for chemical behavior of An-based extended structures, and therefore, shed light on possible optimization of nuclear waste administration.

INTRODUCTION Modularity of hybrid frameworks is an attractive and desirable foundation for development of new constituents, motifs, and architectures for efficient storage, separation, and selective sequestration of nuclear waste, which could address current challenges, especially in light of recently reported problems.1–8 Framework versatility,9–18 in combination with its modularity,19–27 can lead to a more homogeneous actinide distribution (e.g., through actinide metal nodes28–39 or anchoring the actinide to organic linkers40–45), which decreases the accumulation of possible radiation damage caused by the formation of vacancies and defects.46–49 In addition to actinide inclusion, hybrid frameworks also offer the opportunity for actinide immobilization through covalent bond formation. Furthermore, the solvothermal approach commonly used for preparation of metal-organic frameworks (MOFs) relies on moderate temperatures, which prevents formation of volatile radioactive species, in contrast to a ~1000 °C temperature regime required for preparation of radionuclide-containing borosilicate glass as contaminant sequesters.47 In this work, we applied a sequential multi-step approach utilizing MOF modularity and versatility to (i) prepare the first examples of An-bimetallic MOFs through metal node extension and transmetallation, both of which occurred through single-crystal-to-single-crystal transformations; the latter allowed us to demonstrate a drastic difference in the

chemical behavior between molecular species versus extended structures (Scheme 1), (ii) perform the first postsynthetic capping linker installation on An-integrated systems, (iii) demonstrate sequential installation of two capping linkers including one with a selective actinide binding site, (iv) test the possibility of simultaneous capping linker installation and An-containing guest inclusion on bimetallic and monometallic frameworks, (v) prepare a Th-containing framework possessing the largest pore aperture and highest measured surface area known to date, and (vi) synthesize a framework with the highest Th wt% and minimal structural density currently reported. These findings were possible due to synthesis of the first examples of An-containing frameworks with “unsaturated” metal nodes, i.e., the number of organic linkers coordinated to one metal node is less than the maximum possible such as 12.22 The following discussion in this paper is organized by synthetic strategies, which were used to build stepwise hierarchical complexity of An-integrated systems. Comprehensive analysis of materials and their precursors, including single-crystal and powder X-ray diffraction (PXRD), inductively coupled plasma atomic emission spectroscopy (ICP), thermogravimetric and gas sorption analyses, nuclear magnetic resonance (NMR), Fourier transform infrared (FTIR), and X-ray photoelectron spectroscopies (XPS), and theoretical modeling is also discussed for each system separately.

1 ACS Paragon Plus Environment

Journal of the American Chemical Society

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

RESULTS AND DISCUSSION Scheme 1. A Schematic Representation of Precursors (AnMOF and Zr-MOF) and Synthetic Strategies Utilized for Actinide Integration Inside the Rigid Framework. A Set of Organic Linkers Utilized for MOF Preparation and Postsynthetic Capping Linker Installationa

Page 2 of 21

carboxylic groups attached to a metal node), and one Zr-MOF (PCN-700).50 The novel actinide-containing frameworks, Th6O4(OH)6(TFA)2(Me2BPDC)5 (Th6-Me2BPDC-10; Me2BPDC2- = 2,2′-dimethylbiphenyl-4,4′′-dicarboxylate, TFA = trifluoroacetic acid, Figures S1 and S2), U6O4(OH)8(Me2BPDC)4(DMF)2 (U6-Me2BPDC-8, DMF = N,N’-dimethylformamide, Figures S3 and S4), and Th6O4(OH)4(TPDC-NH2)6 (Th6-TPDC-NH2-12; TPDC-NH22= 2′-amino-terphenyl-4,4′′-dicarboxylate, Figures S5 and S6) were prepared using the solvothermal method. Detailed experimental conditions and procedures used for An-MOF synthesis are given in Table 1 and discussed in the Experimental Section (vide infra). Both synthesized Th6-Me2BPDC-10 and U6-Me2BPDC-8 contain “unsaturated” metal nodes An6O4(OH)xLy (An = U, Th; x = 8 (U) or 6 (Th); y = 4 (U) or 5 (Th)) shown in Figure 1. These frameworks represent the first examples of AnMOFs, which could be used as precursors for metal node extension and/or capping linker installation. In contrast, the third example of a synthesized An-MOF, Th6-TPDC-NH2-12, belongs to a series of MOFs possessing UiO-topology (UiO = University of Oslo).51,52 The Th6-TPDC-NH2-12 framework possesses the largest pore aperture (20 × 28 Å) and the highest BET surface area (880 m2/g, Figure S6) reported for Th-based MOFs to date, which opens the possibility to increase actinide content through guest inclusion.

Capping Linker Installation

a

The red color indicates actinide location: red spheres represent An-based metal nodes; grey spheres – Zr-based metal nodes; grey solid sticks – organic linkers used for framework synthesis; blue springs – capping linker; red springs – capping linkers functionalized with an anchoring group; red icosahedra – noncovalently bound actinide-containing guests (UO22+, Th4+).

To build hierarchical complexity of An-based materials stepwise, we prepared four precursors shown in Figure 1: three actinide-containing MOFs, M-Linker-n (M (Th, U) = a metal in the node, Linker = an organic linker, and n = number of

Installation of the capping linker was probed on the example of Th6-Me2BPDC-10 (Table 1), which possesses “unsaturated” metal nodes, and therefore, satisfies the initial criteria necessary to perform this synthetic approach. The choice of capping linkers is based on the size of the pocket between metal nodes in a parent framework, where additional linkers can be installed. For instance, the capping linker, H2TPDCNH2 (length = 15 Å), was chosen because of the Th6Me2BPDC-10 topology to cap cylindrical pores with the 16 × 16 Å channels (Figure 2). In this case, 80% installation of the capping linker was achieved according to 1H NMR spectroscopic analysis by heating Th6-Me2BPDC-10 in a DMF solution of H2TPDC-NH2 at 75 °C for 24 h (Figures S7–S9). The synthesized Th6-Me2BPDC(TPDC-NH2) (here and throughout the manuscript the capping linker is designated in parentheses) is the first example of an actinide-containing MOF successfully utilized for postsynthetic capping linker installation demonstrating

Figure 1. (left) Crystal structures and metal nodes of frameworks utilized as precursors for building hierarchical complexity (from left to right): Zr6-Me2BPDC-8, Th6-Me2BPDC-10, U6-Me2BPDC-8, and Th6-TPDC-NH2-12. (right) Two organic linkers used for framework synthesis are also shown. Red, purple, black, pink, and grey spheres represent Th, U, Zr, O, and C atoms, respectively. Hydrogen atoms and solvents molecules were omitted for clarity.

2 ACS Paragon Plus Environment

Page 3 of 21

Journal of the American Chemical Society

Table 1. Synthetic conditions for An-containing MOFs

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

Compound

Precursors

T,°C/t, h

U6-Me2BPDC-8

UCl4, H2Me2BPDC

120/7

Th6-Me2BPDC-10

ThCl4, H2Me2BPDC

120/9

Th6-TPDC-NH2-12

ThCl4, H2TPDC-NH2

120/72

Th5.65U0.35-Me2BPDC-8

U6-Me2BPDC-8, ThCl4

rt/72

Zr6U0.87-Me2BPDC-8

Zr6-Me2BPDC-8 UO2(CH3COO)2

75/72

Th6U4-Me2BPDC-8

Th6-Me2BPDC-10 UO2(CH3COO)2

75/72

Th6-Me2BPDC (TPDC-NH2)

Th6-Me2BPDC, H2TPDC-NH2

75/24

Zr6-Me2BPDC DEPU)(NDC)

(TPDC-1) Zr6-Me2BPDC, H2TPDC-DEPU

shorter linker, such as NDC2– (NDC2– = naphthalene-2,6dicarboxylate, Scheme 1). For H2TPDC-DEPU installation, crystals of Zr6-Me2BPDC-8 were heated in a DMF solution of the capping linker at 75 °C for 24 h, which resulted in 53% linker installation. In the second step, we performed simultaneous guest inclusion and capping linker installation. For that, Zr6-Me2BPDC(TPDC-DEPU) was heated in a DMF solution of H2NDC (Figure 4) in the presence of uranyl acetate at 75 °C for 24 h, which resulted in 76% installation of the second linker, NDC2–, according to 1H NMR spectroscopic analysis (Figures S10 and S11). The simultaneous installation of the capping linker and incorporation of guest species (UO22+) resulted in a high amount of uranium (44 wt%) immobilized in the pores of the material. After installation of H2TPDC-DEPU as a capping linker, we observed a relatively small amount of uranyl ions coordinated to the anchoring group (95%, International Bio-Analytical Industries Inc.), ThCl4 (>95%, International Bio-Analytical Industries Inc.),

Figure 8. Packing and metal nodes of U6-Me2BPDC-8 and Th6Me2BPDC-8. Insets show the color change occurred during cation exchange process. Red, purple, pink, and grey spheres represent Th, U, O, and C atoms, respectively. Hydrogen atoms and solvents molecules were omitted for clarity. UO2(NO3)2⋅6H2O (98%, International Bio-Analytical Industries Inc.), UO2(CH3COO)2⋅2H2O (98%, International Bio-Analytical Industries Inc.), ZrCl4 (99.5%, Alfa Aesar), CsF (99%, Oakwood Chemical), KOH (ACS grade, Fisher Chemical), K2CO3 (lab grade, Ward’s Science), 2,5-dibromoaniline (97%, Oakwood Chemical), 4methoxycarbonyl phenylboronic acid (>97%, Boronic Molecular), 2,6-naphthalene-dicarboxylic acid (>98%, TCI), stilbene-4,4’dicarboxylic acid (98%, AK Scientific), Pd(OAc)2 (>95%, Ox-Chem), triphenylphosphine (99%, Sigma-Aldrich), diethoxyphosphinyl isocyanate (>90%, Alinda Chemical Ltd.), methyl 4-iodo-3methylbenzoate (98%, BeanTown Chemical), 4,4,4’,4’,5,5,5’,5’octamethyl-2,2’-bi(1,3,2-dioxaborolane) (>98%, Ark Pharm), PdCl2(PPh3)2 (96%, Oakwood Chemical), trifluoroacetic acid (99%, Sigma Aldrich), tetrahydrofuran (ACS grade, Macron Fine Chemicals), dichloromethane (ACS grade, Oakwood Chemical), methanol (ACS grade, Fischer Scientific), diethyl ether (ACS grade, J. T. Baker® Chemicals), dimethyl sulfoxide (ACS grade, Fisher Scientific), N,N’-dimethylformamide (ACS grade, BDH), chloroformd (Cambridge Isotope Laboratories, Inc.), and DMSO-d6 (Cambridge Isotope Laboratories, Inc.) were used as received. Synthesis. The compounds 2,2'-dimethylbiphenyl-4,4'-dicarboxylic acid (H2Me2BPDC),65 2'-amino-[1,1':4',1''-terphenyl]-4,4''dicarboxylic acid (H2TPDC-NH2),66 2'-(3(diethoxyphosphoryl)ureido)-[1,1':4',1''-terphenyl]-4,4''-dicarboxylic acid (H2TPDC-DEPU),40 and Zr6-Me2BPDC-850 were prepared according to the reported procedures. Synthesis and characterization of Th6-Me2BPDC-10. A mixture of ThCl4 (24 mg, 65 µmol), H2Me2BPDC (4.3 mg, 16 µmol), trifluoroacetic acid (25 µL), and DMF (0.75 mL) were mixed in a 1dram vial. The mixture was heated at 120 °C on a hot plate for 9 h. After cooling to room temperature, the colorless crystals of Th6Me2BPDC-10 (6.0 mg, 2.1 µmol, yield: 65%) were collected by filtration and washed three times with DMF. IR (neat, cm-1): 2920, 2853, 1665, 1593, 1546, 1408, 1382, 1255, 1206, 1090, 1006, 910, 863, 777, 732, and 658 (Figure S2). The metal node and packing of Th6-Me2BPDC-10 are shown in Figure 1. The detailed description of the data collection and refinement details are given in the Supporting Information (SI), and Table S1 contains the crystallographic refinement data. As shown in Figure S1, the PXRD pattern of Th6Me2BPDC-10 matches the one simulated from single-crystal X-ray data. Moreover, PXRD studies were used to confirm crystallinity of bulk material (Figure S1). The thermal stability of Th6-Me2BPDC-10 was studied by thermogravimetric analysis, which demonstrated the rapid loss of solvent molecules at the 25–125 °C temperature range

6 ACS Paragon Plus Environment

Page 7 of 21

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

Journal of the American Chemical Society

(Figure S1). The observed weight loss (~32 wt%) at this temperature range can be attributed to removal of the non-coordinated solvent mixture of DMF and H2O, which is in good correlation with the residual electron density calculated from the single-crystal X-ray data. Furthermore, the samples were characterized by FTIR spectroscopy and gas sorption analysis as shown in Figure S2. Before gas sorption analysis, the as-synthesized MOF was washed with DMF and evacuated at 160 °C for 24 h. Fitting the N2 adsorption isotherm to the Brunauer-Emmett-Teller (BET) equation resulted in a surface area of 741 m2/g (Figure S2). Synthesis of and characterization of U6-Me2BPDC-8. A mixture of UCl4 (12 mg, 32 µmol), H2Me2BPDC (4.3 mg, 16 µmol), trifluoroacetic acid (25 µL), and DMF (0.29 mL) were mixed in a 1/2dram vial. The mixture was heated at 120 °C on a hot plate for 7 h. After cooling to room temperature, the green crystals of U6Me2BPDC-8 (7.0 mg, 2.4 µmol, yield: 61%) were retained in the mother liquor. IR (neat, cm-1): 2930, 2862, 1574, 1498, 1409, 1378, 1253, 1195, 1090, 1062, 1006, 916, 865, 780, 673, and 657 (Figure S4). The obtained crystals were suitable for single-crystal X-ray analysis. The metal node and packing of U6-Me2BPDC-8 are shown in Figure 1. The detailed description for the data collection and refinement details are given in SI. Table S1 contains the crystallographic refinement data. As shown in Figure S3, the PXRD pattern of U6-Me2BPDC-8 matches the one simulated from singlecrystal X-ray data. PXRD studies were also used to confirm crystallinity of bulk material (Figure S3). Thermal stability of U6Me2BPDC-8 was studied by thermogravimetric analysis (Figure S3), which demonstrated a rapid weight loss (~ 38 wt%) occurring at the 25–300 °C temperature range. Synthesis and characterization of Th6-TPDC-NH2-12. A mixture of ThCl4 (32 mg, 86 µmol), H2TPDC-NH2 (20 mg, 60 µmol), acetic acid (100 µL), and DMF (4 mL) were mixed in a 2-dram vial. The mixture was heated at 120 °C in an oven for 72 h. After cooling, the crystals of Th6-TPDC-NH2-12 (25 mg, 6.9 µmol, yield: 69%) were collected by filtration and washed three times with DMF. IR (neat, cm-1): 3342, 2928, 1659, 1597, 1549, 1386, 1253, 1180, 1089, 864, 838, 780, 709, and 658 (Figure S6). The metal node and packing of Th6-Me2BPDC-12 are shown in Figure 1. The detailed description for the data collection and refinement details are given in the SI, and Table S1 contains the crystallographic refinement data. As shown in Figure S5, the PXRD pattern of Th6-TPDC-NH2-12 matches the one simulated based on the single-crystal X-ray analysis. PXRD studies were used to confirm crystallinity of bulk material (Figure S5). Thermal stability of Th6-Me2BPDC-12 was studied by thermogravimetric analysis (Figure S5). The observed weight loss (~42 wt%) in could be attributed to removal of the non-coordinated solvent mixture of DMF and H2O used for MOF synthesis, which is in good correlation with the residual electron density calculated from single-crystal X-ray data. Furthermore, the samples were characterized by FTIR and gas sorption analysis as shown in Figure S6. Before gas sorption analysis, the as-synthesized MOF was washed with DMF and evacuated at 200 °C for 10 h. Fitting the N2 adsorption isotherm to the BET equation resulted in a surface area of 880 m2/g (Figure S6). Synthesis and characterization Zr6U0.87-Me2BPDC-8. To perform metal node extension, 25 mg of Zr6-Me2BPDC-8 (PCN700)50 was soaked in 1.0 mL of the 0.25 M uranyl acetate solution in DMF, and then kept in a pre-heated oven at 75 °C for 3 d. Yellow crystals of Zr6U0.87-Me2BPDC-8 were collected by filtration and washed three times with DMF. The metal node and crystal structure of Zr6U0.87-Me2BPDC-8 are shown in Figure 5. The detailed description for the data collection and refinement details are given in SI, and Table S3 contains the crystallographic refinement data. As shown in Figure S12, the PXRD pattern of Zr6U0.87-Me2BPDC-8 matches the one simulated from the single-crystal X-ray data. Therefore, PXRD studies demonstrate preservation of framework integrity after the metal node extension (Figure S12). Before ICPAES analysis the prepared sample was thoroughly washed using a Soxhlet extraction for 3 d to remove possible residual salt. The Zr-toU metal ratio was determined based on the ICP-AES analysis. FTIR

spectroscopy was employed to confirm presence of uranyl unit due to the presence of the U=O vibrational stretches at 913 and 866 cm-1 (Figure S13). Thermal stability of Zr6U0.87-Me2BPDC-8 was studied using thermogravimetric analysis (Figure S12), which demonstrated a rapid weight loss occurring up to 320 °C. XPS studies demonstrate presence of Zr+4 and U+6 in the sample, which is consistent with the single-crystal X-ray studies and FTIR spectroscopic data (Figures 5 and S14). Synthesis and characterization of Th6U4-Me2BPDC-8. To achieve metal node extension, 25 mg of Th6-Me2BPDC-10 was soaked in 1.0 mL of 0.25 M uranyl acetate solution in DMF and kept in a pre-heated oven at 75 °C for 3 d. The obtained yellow crystals were collected by filtration and washed three times with fresh DMF, and were subjected to single-crystal X-ray analysis (Figure 6). Table S3 contains the crystallographic refinement data. As shown in Figure S17, the PXRD pattern of Th6U4-Me2BPDC-8 matches one simulated from single-crystal X-ray data, and therefore, MOF integrity was preserved after metal node extension (Figure S15). The Th-to-U ratio was determined based on the ICP-AES analysis, but before the ICPAES analysis, the prepared sample was thoroughly washed using a Soxhlet extraction for 3 d to remove residual metal salts. Thermogravimetric analysis was used to study thermal stability of the presented samples (Figure S15), which demonstrated a rapid weight loss up to 300 °C. XPS studies confirmed presence of U+6 and Th+4 in the sample, which is consistent with the Th-node extension determined based on single-crystal X-ray analysis (Figure S17). Transmetallation attempts for Zr6-Me2BPDC-8. To explore the possibility of Zr-to-Th transmetallation, the colorless crystals of Zr6Me2BPDC-8 (20 mg, 11 µmol) were heated in 2.0 mL of ThCl4 solution (C = 67 mM) in DMF at 75 °C for 14 d. The obtained sample was characterized by single-crystal X-ray diffraction and spectroscopic studies, both of which did not reveal the presence of thorium in the MOF skeleton and in line with our estimated energy of formation (Table S4). Synthesis and characterization of Th5.65U0.35-Me2BPDC-8. The green crystals of U6-Me2BPDC-8 were washed once with DMF and soaked in a 0.20 mL solution of 0.17 M ThCl4 in DMF for 3 d at room temperature. After 3 d, the resulting colorless crystals of Th5.65U0.35Me2BPDC-8 were collected by centrifugation and washed thoroughly three times with DMF. The obtained colorless crystals were suitable for single-crystal X-ray analysis (Figure 8). Table S4 contains the crystallographic refinement data. As shown in Figure S20, the PXRD pattern of Th5.65U0.35-Me2BPDC-8 matches the one simulated from the single-crystal X-ray data. PXRD studies were used to confirm crystallinity of bulk material (Figure S20). The Th-to-U ratio was determined based on the ICP-AES analysis. Thermal stability of Th5.65U0.35-Me2BPDC-8 was studied using thermogravimetric analysis (Figure S20). The FTIR spectrum of Th5.65U0.35-Me2BPDC-8 is shown in Figure S21. General Procedure of Capping linker installation. Compounds Th6-Me2BPDC(TPDC-NH2), Zr6-Me2BPDC(TPDC-DEPU)(NDC) (NDC2= naphthalene-2,6-dicarboxylate), Th5.65U0.35Me2BPDC(SDC) (SDC2- = 4,4’-stilbenecarboxylate), and Zr6U0.87Me2BPDC(SDC) were synthesized by the capping linker installation in parent Th6-Me2BPDC-10, Zr6-Me2BPDC-8,50 Th5.65U0.35Me2BPDC-8, and Zr6U0.87-Me2BPDC-8, respectively, based on the modified literature procedure.50 Crystals of the parent MOFs were heated in a DMF solution of a corresponding capping linker at 75 °C for 24 h (Table 1). The obtained crystals were collected by filtration. The washing procedure includes thoroughly washing with hot DMF to remove the residual capping linker. Simultaneous incorporation of actinides as guests and capping linker installation was performed by heating UO2(CH3COO)2 or ThCl4, the parent MOF, and the organic linker of interest in DMF at 75 °C for 24 h. Digestion procedure. To study the composition of the prepared MOFs by 1H NMR spectroscopy, a solution of 500 µL of DMSO-d6 and 3 µL of concentrated HCl was added to 5 mg of the material, followed by sonication until complete sample dissolution. The % of capping linker installation was calculated based on linker ratios found in the 1H NMR spectra of the digested samples. The highest amount

7 ACS Paragon Plus Environment

Journal of the American Chemical Society

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

of capping linker can be calculated from geometrical analysis of parent MOF structure and length of capping linker as shown by Zhou and co-workers.53 Synthesis and characterization of Th6-Me2BPDC(TPDC-NH2). The Th6-Me2BPDC(TPDC-NH2) framework was synthesized through installation of the capping linker, H2TPDC-NH2, into parent Th6Me2BPDC-10. The crystals of Th6-Me2BPDC-10 (20.0 mg, 6.28 µmol) were heated in 4 mL of 15.0 mM H2TPDC-NH2 solution in DMF at 75 °C in a pre-heated oven for 24 h. To remove the residual capping linker, as-synthesized Th6-Me2BPDC(TPDC-NH2) was thoroughly washed with hot DMF. Based on 1H NMR spectroscopic analysis of the digested sample (Figure S7), we found that 80% of the capping linker, TPDC-NH22–, was installed. The installation of TPDC-NH22– was also confirmed by the presence of corresponding to –NH and –CN stretches in the FTIR spectrum of Th6Me2BPDC(TPDC-NH2) (Figure S8). PXRD studies were used to confirm preservation of framework integrity after capping linker installation (Figure S9). Synthesis and characterization of Zr6-Me2BPDC(TPDCDEPU)(NDC). The Zr6-Me2BPDC(TPDC-DEPU)(NDC) framework was synthesized through stepwise installation of two different capping linkers into parent Zr6-Me2BPDC-8. In the first step, crystals of Zr6Me2BPDC-8 (15.0 mg, 7.96 µmol) were heated in 1 mL of H2TPDCDEPU (C = 30.0 mM) solution in DMF at 75 °C for 24 h. The obtained single crystals thoroughly washed with DMF were still suitable for single-crystal X-ray diffraction. The detailed description for the data collection and refinement details are given in Table S2. In the second step, obtained Zr6-Me2BPDC(TPDC-DEPU) was heated in 1 mL of H2NDC solution (C = 30.0 mM) in DMF at 75 °C for 24 h. The resulting Zr6-Me2BPDC(TPDC-DEPU)(NDC) framework was thoroughly washed with hot DMF to remove the residual capping linker(s). The composition of Zr6-Me2BPDC(TPDC-DEPU)(NDC) was determined based on 1H NMR spectroscopy (Figure S10). The installation of H2TPDC-DEPU and H2NDC was found to be 53% and 76%, respectively. PXRD studies were used to confirm MOF integrity after capping linker installation (Figure S11). Synthesis and characterization of Zr6U0.87-Me2BPDC(SDC). The Zr6U0.87-Me2BPDC(SDC) framework was synthesized through the installation of the capping linker, SDC2–, into parent Zr6U0.87Me2BPDC-8. The crystals of Zr6U0.87-Me2BPDC-8 (25.0 mg, 11.3 µmol) were heated in the 10 mL solution of H2SDC (C = 8.0 mM) in DMF at 75 °C for 24 h. The resulting solid was collected by centrifugation and washed thoroughly with hot DMF. Based on 1H NMR spectroscopic studies (Figure S18), the SDC2– capping linker installation was found to be 50%. PXRD studies were used to confirm MOF integrity after capping linker installation (Figure S19). Synthesis and characterization of Th5.65U0.35-Me2BPDC(SDC). The Th5.65U0.35-Me2BPDC(SDC) framework was synthesized through installation of a capping linker, H2SDC, into parent Th5.65U0.35Me2BPDC-8. The crystals of Th5.65U0.35-Me2BPDC-8 (7.00 mg, 2.54 µmol) were soaked in a DMF solution of H2SDC (8.00 mM, 4 mL) at 75 °C for 24 h. The resulting solid was collected by centrifugation, washed thoroughly with hot DMF. Based on 1H NMR spectroscopic analysis of the digested framework, SDC2– installation was found to be 76% (Figure S23). PXRD analysis confirmed preservation of MOF integrity after capping linker installation. As expected due to additional coordination of the capping linker, SDC2–, the PXRD pattern of Th5.65U0.35-Me2BPDC(SDC) matches the one of Th6Me2BPDC-10 (Figure S24). Physical measurements. FTIR spectra were obtained on a PerkinElmer Spectrum 100. NMR spectra were collected on Bruker Avance III-HD 300 and Bruker Avance III 400 MHz NMR spectrometers. The 1H and 13C NMR spectra were referenced to natural abundance 13 C signals and residual 1H signals of deuterated solvents, respectively. Powder X-ray diffraction patterns were recorded on a Rigaku Miniflex II diffractometer with accelerating voltage and current of 30 kV and 15 mA, respectively. Thermogravimetric analysis was performed on an SDT Q600 Thermogravimetric Analyzer using an alumina boat as a sample holder at a heating rate of 5 °C/min. ICP-AES analysis was conducted using a Finnigan

Page 8 of 21

ELEMENT XR double focusing magnetic sector field inductively coupled plasma-mass spectrometer (SF-ICP-MS) with Ir and/or Rh as internal standards. A Micromist U-series nebulizer (0.2 ml/min, GE, Australia), quartz torch, and injector (Thermo Fisher Scientific, USA) were used for sample introduction. Sample gas flow was 1.08 mL/min, and the forwarding power was 1250 W. The samples were digested in Teflon vessels with nitric and hydrochloric acids and then heated at 180 °C for 4 h. Gas sorption measurements were conducted on a Micromeritics ASAP 2020 system. Oven-dried sample tubes equipped with a TranSeal™ (Micrometrics) were evacuated and tarred. Samples were transferred to the sample tube, which were then capped by a TranSeal™. Samples were heated to the appropriate temperatures as determined by TGA. The evacuated sample tubes were weighed again and the sample mass was determined by subtracting the mass of the previously tarred tube. N2 isotherms were measured using a liquid nitrogen bath (77 K). Ultra-high purity grade (99.999% purity) N2 and He, oil-free valves and gas regulators were used for all free space corrections and measurements. X-ray photoelectron spectroscopy measurements (XPS) were performed using a Kratos AXIS Ultra DLD XPS system with a monochromatic Al Kα source operated at 15 keV and 150W and a hemispherical energy analyzer. Samples were placed in small powder pockets on the holder and analysis was performed at a pressure below 1×10–9 mbar. High-resolution core level spectra were measured with a pass energy of 40 eV, and analysis of the data was carried out using XPSPEAK41 software. The XPS experiments were performed while using an electron gun directed on the sample, for charge neutralization. Theoretical calculations. Calculations were performed using the VASP software67,68 with plane wave basis sets. Projector-augmented-wave (PAW) pseudopotentials69,70 of Zr, Th, U, C, O, and H were employed in which the number of electrons treated as valence is 12, 12, 14, 4, 6, and 1, respectively. The PAW potentials were taken from the VASP library. Calculations were performed with a plane wave energy cutoff of 520 eV and Γ-only k-point. Structure optimization was performed until the norm of the atomic forces is less than 0.025 eV/Å. Two different levels of theory were explored. One was within a pure density-functional-theory with the Perdew-Burke-Erzernhof exchange-correlation functional71 and with a Van der Waals dispersion correction (denoted as PBE-D3). The other was within a hybrid Hartree-Fock/DFT with the B3LYP hybrid functional72 and also with a dispersion correction (denoted as B3LYP-D3). The Van der Waals interactions were taken into account using the dispersion formula of Grimme et al.73 with Becke-Johnson damping.74 In addition, in the PBE-D3 set, an on-site Coulomb interaction was added to the uranium f electrons to improve the electronic structure of these localized electrons within the DFT+U formalism.75 Based on previous studies,76–82 a U-J = 4.0 eV was used. First, we optimized the Zr6(HCO2)8O8 structure. Table 2 shows the relaxed structure obtained with B3LYP-D3, however the structure obtained with PBE-D3 is nearly identical. Subsequently a Th or U atom was substituted for one of the Zr atoms. For this one-atom substitution, there are two unique sites, pos1 (1) and pos3 (3) as shown in Table2. The formation energy for this one-atom substitution (Ef{1}) is calculated from the following total energies (equation 1): Ef{1}=Et{MZr5(HCO2)8O8+Et{Zr4+}–Et{Zr6(HCO2)8O8}-Et{M4+} Where Et{M4+} is the total energy of an isolated positive ion (M = Th or U). The image charge correction due to periodic boundaries has been taken into account by using the static dielectric constant of DMF at room temperature ε = 37.65.83 Furthermore, substitution of all the Zr atoms with Th atoms or U atoms was also investigated. Similarly to Equation 1, the formation energy for this six-atom substitution (Ef{6}) is calculated from: Ef{6}=Et{M6(HCO2)8O8}+6Et{Zr4+}–Et{Zr6(HCO2)8O8}–6Et{M4+} Table 2 summarizes the formation energies. The results show that substitutions with Th or U are not energetically favored.

8 ACS Paragon Plus Environment

Page 9 of 21

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

Journal of the American Chemical Society

ASSOCIATED CONTENT Supporting Information The Supporting Information is available free of charge on the ACS Publications website. X-ray structure refinement data, packing and metal nodes of prepared MOFs, PXRD patterns, TGA traces, FTIR, XPS, and 1H NMR spectra (PDF) Crystallographic information for Th6-TPDC-NH2-12, CCDC 1551809 (CIF) Crystallographic information for Th5.65U0.35-Me2BPDC-8, CCDC 1551811 (CIF) Crystallographic information for Zr6U0.87-Me2BPDC-8, CCDC 1551812 (CIF) Crystallographic information for Th6-Me2BPDC-10, CCDC 1551813 (CIF) Crystallographic information for Th6U4-Me2BPDC-8, CCDC 1552322 (CIF) Crystallographic information for U6-Me2BPDC-8, CCDC 1552208 (CIF) Crystallographic information for PCN-700(Th), CCDC 1552392 (CIF) Crystallographic information for Zr6-Me2BPDC(TPDC-DEPU), CCDC 1580371 (CIF)

AUTHOR INFORMATION Corresponding Author * [email protected]

Notes The authors declare no competing financial interest.

ACKNOWLEDGMENT This research was supported as part of the Center for Hierarchical Wasteform Materials (CHWM), an Energy Frontier Research Center funded by the U.S. Department of Energy, Office of Science under Award DE-SC0016574. N.B.S. acknowledges the support from the Sloan Research Fellowship provided by Alfred P. Sloan Foundation. E.A.D. acknowledges support of the Office of the Vice President for Research in the form of a SPARC Graduate Research Grant.

REFERENCES (1) Morrison, J. C&EN Glob. Enterp. 2017, 95, 26–27. (2) Edelmann, F. T. Coord. Chem. Rev. 2017, 338, 27–140. (3) Council, N. R. Waste Forms Technology and Performance; National Academies Press: Washington, D.C., 2011. (4) Ewing, R. C.; von Hippel, F. N. Science 2009, 325, 151–152. (5) Weber, W. J.; Navrotsky, A.; Stefanovsky, S.; Vance, E. R.; Vernaz, E. MRS Bull. 2009, 34, 46–53. (6) Walther, C.; Denecke, M. A. Chem. Rev. 2013, 113, 995–1015. (7) Natrajan, L. S.; Swinburne, A. N.; Andrews, M. B.; Randall, S.; Heath, S. L. Coord. Chem. Rev. 2014, 266–267, 171–193. (8) Li, D.; Egodawatte, S.; Kaplan, D. I.; Larsen, S. C.; Serkiz, S. M.; Seaman, J. C. J. Hazard. Mater. 2016, 317, 494–502. (9) Yaghi, O. M.; O’Keeffe, M.; Ockwig, N. W.; Chae, H. K.; Eddaoudi, M.; Kim, J. Nature 2003, 423, 705–714. (10) Schoedel, A.; Li, M.; Li, D.; O’Keeffe, M.; Yaghi, O. M. Chem. Rev. 2016, 116, 12466–12535. (11) Furukawa, H.; Cordova, K. E.; O’Keeffe, M.; Yaghi, O. M. Science 2013, 341, 1230444.

(12) Cohen, S. M. J. Am. Chem. Soc. 2017, 139, 2855–2863. (13) Brozek, C. K.; Dincă, M. Chem. Soc. Rev. 2014, 43, 5456– 5467. (14) Zhang, T.; Lin, W. Chem. Soc. Rev. 2014, 43, 5982–5993. (15) Huang, F.-P.; Yao, P.-F.; Li, H.-Y.; Yu, Q.; Bian, H.-D.; Liang. H. Chem. Commun. 2015, 51, 7598–7601. (16) Gallis, D. F. S.; Ermanoski, I.; Greathouse, J. A.; Chapman, K. W.; Nenoff, T. M. Ind. Eng. Chem. Res. 2017, 56, 2331–2338. (17) Demir, S.; Brune, N. K.; Van Humbeck, J. F.; Mason, J. A.; Plakhova, T. V.; Wang, S.; Tian, G.; Minasian, S. G.; Tyliszczak, T.; Yaita, T.; Kobayashi, T.; Kalmykov, S. N.; Shiwaku, H.; Shuh, D. K.; Long, J. R. ACS Cent. Sci. 2016, 2, 253–265. (18) Xie, J.; Wang, Y.; Liu, W.; Yin, X.; Chen, L.; Zou, Y.; Diwu, J.; Chai, Z.; Albrecht-Schmitt, T. E.; Liu, G.; Wang, S. Angew. Chem. Int. Ed. 2017, 56, 7500–7504. (19) Odoh, S. O.; Shamblin, J.; Colla, C. A.; Hickam, S.; Lobeck, H. L.; Lopez, R. A. K.; Olds, T.; Szymanowski, J. E. S.; Sigmon, G. E.; Neuefeind, J.; Casey, W. H.; Lang, M.; Gagliardi, L.; Burns, P. C. Inorg. Chem. 2016, 55, 3541–3546. (20) Lussier, A. J.; Lopez, R. A. K.; Burns, P. C. Can. Mineral. 2017, 54, 177–283. (21) Yang, H.; Luo, M.; Luo, L.; Wang, H.; Hu, D.; Lin, J.; Wang, X.; Wang, Y.; Wang, S.; Bu, X.; Feng, P.; Wu, T. Chem. Mater. 2016, 28, 8774–8780. (22) Rimoldi, M.; Howarth, A. J.; DeStefano, M. R.; Lin, L.; Goswami, S.; Li, P.; Hupp, J. T.; Farha, O. K. ACS Catal. 2017, 7, 997–1014. (23) Ullman, A. M.; Brown, J. W.; Foster, M. E.; Léonard, F.; Leong, K.; Stavila, V.; Allendorf, M. D. Inorg. Chem. 2016, 55, 7233–7249. (24) Lalonde, M.; Bury, W.; Karagiaridi, O.; Brown, Z.; Hupp, J. T.; Farha, O. K. J. Mater. Chem. A 2013, 1, 5453–5468. (25) Gardinier, J. R.; Hewage, J. S.; Hoffman, J.; Lindeman, S. V.; Williams, D. E.; Shustova, N. B. Eur. J. Inorg. Chem. 2016, 2016, 2615–2625. (26) Baudron, S. A. CrystEngComm 2016, 18, 4671–4680. (27) Béziau, A.; Baudron, S. A.; Rogez, G.; Hosseini, M. W. Inorg. Chem. 2015, 54, 2032–2039. (28) Li, P.; Vermeulen, N. A.; Malliakas, C. D.; Gómez-Gualdrón, D. A.; Howarth, A. J.; Mehdi, B. L.; Dohnalkova, A.; Browning, N. D.; O’Keeffe, M.; Farha, O. K. Science 2017, 356, 624–627. (29) Li, P.; Vermeulen, N. A.; Gong, X.; Malliakas, C. D.; Stoddart, J. F.; Hupp, J. T.; Farha, O. K. Angew. Chem. Int. Ed. 2016, 55, 10358–10362. (30) Andrews, M. B.; Cahill, C. L. Chem. Rev. 2013, 113, 1121– 1136. (31) Ok, K. M.; Sung, J.; Hu, G.; Jacobs, R. M. J.; O’Hare, D. J. Am. Chem. Soc. 2008, 130, 3762–3763. (32) Falaise, C.; Charles, J.-S.; Volkringer, C.; Loiseau, T. Inorg. Chem. 2015, 54, 2235–2242. (33) Falaise, C.; Volkringer, C.; Vigier, J.-F.; Henry, N.; Beaurain, A.; Loiseau, T. Chem. Eur. J. 2013, 19, 5324–5331. (34) Kalaj, M.; Carter, K. P.; Savchenkov, A. V.; Pyrch, M. M.; Cahill, C. L. Inorg. Chem. 2017, 56, 9156–9168. (35) Martin, N. P.; März, J.; Volkringer, C.; Henry, N.; Hennig, C.; Ikeda-Ohno, A.; Loiseau, T. Inorg. Chem. 2017, 56, 2902–2913. (36) Bai, Z.; Wang, Y.; Li, Y.; Liu, W.; Chen, L.; Sheng, D.; Diwu, J.; Chai, Z.; Albrecht-Schmitt, T. E.; Wang, S. Inorg. Chem. 2016, 55, 6358–6360. (37) Berg, J. M.; Gaunt, A. J.; May, I.; Pugmire, A. L.; Reilly, S. D.; Scott, B. L.; Wilkerson, M. P. Inorg. Chem. 2015, 54, 4192–4199. (38) Knope, K. E.; Soderholm, L. Chem. Rev. 2013, 113, 944–994. (39) Tutson, C. D.; Gorden, A. E. V. Coord. Chem. Rev. 2017, 333, 27–43. (40) Carboni, M.; Abney, C. W.; Liu, S.; Lin, W. Chem. Sci. 2013, 4, 2396–2402. (41) Charlot, A.; Dumas, T.; Solari, P. L.; Cuer, F.; Grandjean, A. Eur. J. Inorg. Chem. 2017, 2017, 563–573. (42) Leybros, A.; Hung, L.; Hertz, A.; Hartmann, D.; Grandjean, A.; Boutin, O. Chem. Eng. J. 2017, 316, 196–203.

9 ACS Paragon Plus Environment

Journal of the American Chemical Society

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

(43) Li, B.; Dong, X.; Wang, H.; Ma, D.; Tan, K.; Shi, Z.; Chabal, Y. J.; Han, Y.; Li, J. Faraday Discuss. 2017, 109, 3499–3506. (44) Li, L.; Ma, W.; Shen, S.; Huang, H.; Bai, Y.; Liu, H. ACS Appl. Mater. Interfaces 2016, 8, 31032–31041. (45) Abney, C. W.; Das, S.; Mayes, R. T.; Kuo, L.-J.; Wood, J.; Gill, G.; Piechowicz, M.; Lin, Z.; Lin, W.; Dai, S. Phys. Chem. Chem. Phys. 2016, 18, 23462–23468. (46) Ewing, R. C.; Weber, W. J.; Clinard, F. W. Prog. Nucl. Energy 1995, 29, 63–127. (47) Weber, W. J.; Ewing, R. C.; Catlow, C. R. A.; de la Rubia, T. D.; Hobbs, L. W.; Kinoshita, C.; Matzke, H.; Motta, A. T.; Nastasi, M.; Salje, E. K. H.; Vance, E. R.; Zinkle, S. J. J. Mater. Res. 1998, 13, 1434–1484. (48) Amoroso, J.; Marra, J.; Conradson, S. D.; Tang, M.; Brinkman, K. J. Alloys Compd. 2014, 584, 590–599. (49) Clark, B. M.; Tumurgoti, P.; Sundaram, S. K.; Amoroso, J. W.; Marra, J. C.; Shutthanandan, V.; Tang, M. J. Nucl. Mater. 2017, 494, 61–66. (50) Yuan, S.; Lu, W.; Chen, Y.-P.; Zhang, Q.; Liu, T.-F.; Feng, D.; Wang, X.; Qin, J.; Zhou, H.-C. J. Am. Chem. Soc. 2015, 137, 3177–3180. (51) Cavka, J. H.; Jakobsen, S.; Olsbye, U.; Guillou, N.; Lamberti, C.; Bordiga, S.; Lillerud, K. P. J. Am. Chem. Soc. 2008, 130, 13850– 13851. (52) Schaate, A.; Roy, P.; Godt, A.; Lippke, J.; Waltz, F.; Wiebcke, M.; Behrens, P. Chem. Eur. J. 2011, 17, 6643–6651. (53) Yuan, S.; Chen, Y.-P.; Qin, J.-S.; Lu, W.; Zou, L.; Zhang, Q.; Wang, X.; Sun, X.; Zhou, H.-C. J. Am. Chem. Soc. 2016, 138, 8912– 8919. (54) Qin, J.-S.; Yuan, S.; Alsalme, A.; Zhou, H.-C. ACS Appl. Mater. Interfaces 2017, 9, 33408–33412. (55) Yuan, S.; Zou, L.; Li, H.; Chen, Y.-P.; Qin, J.; Zhang, Q.; Lu, W.; Hall, M. B.; Zhou, H.-C. Angew. Chem. Int. Ed. 2016, 55, 10776– 10780. (56) Yuan, S.; Chen, Y.-P.; Qin, J.; Lu, W.; Wang, X.; Zhang, Q.; Bosch, M.; Liu, T.-F.; Lian, X.; Zhou, H.-C. Angew. Chem. Int. Ed. 2015, 54, 14696–14700. (57) Kim, J.-Y.; Norquist, A. J.; O’Hare, D. J. Am. Chem. Soc. 2003, 125, 12688–12689. (58) Zhang, Y.; Kadi, F.; Karatchevtseva, I.; Price, J. R.; Murphy, T.; Wuhrer, R.; Li, F.; Lumpkin, G. R. J. Incl. Phenom. Macrocycl. Chem. 2015, 82, 163–172. (59) Frisch, M.; Cahill, C. L. Cryst. Growth Des. 2008, 8, 2921– 2928. (60) Li, Y.; Weng, Z.; Wang, Y.; Chen, L.; Sheng, D.; Liu, Y.; Diwu, J.; Chai, Z.; Albrecht-Schmitt, T. E.; Wang, S. Dalt. Trans. 2015, 44, 20867–20873. (61) Li, Y.; Weng, Z.; Wang, Y.; Chen, L.; Sheng, D.; Diwu, J.; Chai, Z.; Albrecht-Schmitt, T. E.; Wang, S. Dalt. Trans. 2016, 45, 918–921. (62) Das, R. K.; Barnea, E.; Andrea, T.; Kapon, M.; Fridman, N.; Botoshansky, M.; Eisen, M. S. Organometallics 2015, 34, 742–752. (63) Brozek, C. K.; Dincă, M. Chem. Sci. 2012, 3, 2110–2113. (64) Brozek, C. K.; Dincă, M. J. Am. Chem. Soc. 2013, 135, 12886–12891. (65) Burrows, A. D.; Frost, C. G.; Mahon, M. F.; Richardson, C. Chem. Commun. 2009, 4218–4220. (66) Manna, K.; Zhang, T.; Carboni, M.; Abney, C. W.; Lin, W. J. Am. Chem. Soc. 2014, 136, 13182–13185. (67) Kresse, G.; Hafner, J. Phys. Rev. B 1993, 47, 558–561. (68) Kresse, G.; Furthmüller, J. Phys. Rev. B 1996, 54, 11169– 11186. (69) Blöchl, P. E. Phys. Rev. B 1994, 50, 17953–17979. (70) Kresse, G.; Joubert, D. Phys. Rev. B 1999, 59, 1758–1775. (71) Perdew, J. P.; Burke, K.; Ernzerhof, M. Phys. Rev. Lett. 1996, 77, 3865–3868. (72) Becke, A. D. J. Chem. Phys. 1993, 98, 5648–5652. (73) Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. J. Chem. Phys. 2010, 132, 154104. (74) Grimme, S.; Ehrlich, S.; Goerigk, L. J. Comput. Chem. 2011, 32, 1456–1465.

Page 10 of 21

(75) Dudarev, S. L.; Botton, G. A.; Savrasov, S. Y.; Humphreys, C. J.; Sutton, A. P. Phys. Rev. B 1998, 57, 1505–1509. (76) Dorado, B.; Amadon, B.; Freyss, M.; Bertolus, M. Phys. Rev. B 2009, 79, 235125–235132. (77) Dudarev, S. L.; Castell, M. R.; Botton, G. A.; Savrasov, S. Y.; Muggelberg, C.; Briggs, G. A. D.; Sutton, A. P.; Goddard, D. T. Micron 2000, 31 (4), 363–372. (78) Gryaznov, D.; Heifets, E.; Sedmidubsky, D. Phys. Chem. Chem. Phys. 2010, 12, 12273–12278. (79) Gupta, F.; Brillant, G.; Pasturel, A. Philos. Mag. 2007, 87, 2561–2569. (80) Zhou, F.; Ozolins, V. Phys. Rev. B 2011, 83, 85106–85110. (81) Wang, B.-T.; Zhang, P.; Lizárraga, R.; Di Marco, I.; Eriksson, O. Phys. Rev. B 2013, 88, 104107–104116. (82) Ramanantoanina, H.; Kuri, G.; Daul, C.; Bertsch, J. Phys. Chem. Chem. Phys. 2016, 18, 19020–19031. (83) Wohlfarth, C. Static Dielectric Constants of Pure Liquids and Binary Liquid Mixtures, Springer, 1991.

10 ACS Paragon Plus Environment

Page 11 of 21

Journal of the American Chemical Society

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

11 ACS Paragon Plus Environment

Journal of the American Chemical Society

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

Figure 1. Crystal structures and metal nodes of frameworks utilized as precursors for building hierarchical complexity (from left to right): Zr6-Me2BPDC-8, Th6-Me2BPDC-10, U6-Me2BPDC-8, and Th6-TPDC-NH2-12. Two organic linkers used for framework synthesis are also shown. Red, purple, black, pink, and grey spheres represent Th, U, Zr, O, and C atoms, respectively. Hydrogen atoms and solvents molecules were omitted for clarity. 37x8mm (600 x 600 DPI)

ACS Paragon Plus Environment

Page 12 of 21

Page 13 of 21

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

Journal of the American Chemical Society

Figure 2. Installation of H2TPDC-NH2 capping linker (blue spring) in Th6-Me2BPDC-10 through coordination to “unsaturated” metal nodes leading to formation of Th6-Me2BPDC(TPDC-NH2). Hydrogen atoms are omitted for clarity. 58x21mm (600 x 600 DPI)

ACS Paragon Plus Environment

Journal of the American Chemical Society

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

Figure 3. Wt% of thorium in MOFs as a function of structural 1/d (d = density*). Red diamonds are this work, blue diamonds – literature data. 1: Th6-Me2BPDC-10, 2: Th5.65U0.35-Me2BPDC(SDC), 3: Th6Me2BPDC-12, 4: Th6O4(OH)4(H2O)6(BDC)6·6DMF ·12H2O,32 5: Th[(BTC)F]0.3H2O,31 6: [(Th2F5)-(3,5-PDC)2 (H2O)][NO3],57 7: Th(BDC)2,32,58 8: Th(2,4-PYDC)2(H2O),59 9: Th(2,3-PYDC)2(H2O)2·2H2O,59 10: Th(2,5-PZDC)2(H2O)2 ·2H2O,59 11: Th(2,3-PZDC)2(H2O)3·H2O,59 12: Th(BTCA)(DMF)2(H2O),58 13: Th(BDC)2(DMF)232,58, 14: Th(2,5-PYDC)(H(2,5-PYDC))2(H2O)3·2H2O,59 15: [Th(TPO)(OH) (H2O)]·8H2O,60 16: [AMIM]2[Th(BTB)Cl3],61 17: [AMIM]5 [Th2(BTB)2Cl6]·Cl,61 18: [BMIM][Th(TPO)Cl2]·18H2O60, 19: [DMA][Th2(NTB)3(H2O)2]·8H2O·6DMF61; (BDC2– = terephthalate, BTC3– = benzene-1,3,5-tricarboxylate, 3,5-PDC2– = 3,5-pyridinedicarboxylate, 2,4-PYDC2– = 2,4-pyridinedicarboxylate, 2,3-PYDC2– = 2,3-pyri-dinedicarboxylate, 2,5-PZDC2– = 2,5pyrazinedicarboxylate, 2,3-PZDC2– = 2,3-pyrazinedicarboxylate, BTCA4– = 1,2,4,5-benzenetetracarboxylate, 2,5-PYDC2– = 2,5-pyridinedicarboxylate, TPO3– = 4,4',4''-(oxo-5phosphanetriyl)tribenzoate, AMIM+ = 1-allyl-3-methylimidazolium, BTB3-= benzene-1,3,5-tribenzoate, BMIM+ = 1-butyl-2, 3-dimethylimida-zolium, DMA = dimethyl amine, NTB3– = 4,4',4''-nitrilo-tribenzoate). *d – structural density determined from single-crystal X-ray data; the low value of d is highlighting MOF porous nature, which is necessary for incorporation of actinide-containing guest species.

144x96mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 14 of 21

Page 15 of 21

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

Journal of the American Chemical Society

Figure 4. Stepwise installation of two capping linkers (red and blue springs) in Zr6-Me2BPDC-8 leading to formation of Zr6-Me2BPDC(TPDC-DEPU)(NDC). 97x56mm (600 x 600 DPI)

ACS Paragon Plus Environment

Journal of the American Chemical Society

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

Figure 5. (top) Packing and metal nodes of Zr6-Me2BPDC-8 and Zr6U0.87-Me2BPDC-8. Insets show photographs of Zr6-Me2BPDC-8 and Zr6U0.87-Me2BPDC-8 powders highlighting the drastic color change during metal node extension. Black, purple, pink, and grey spheres represent Zr, U, O, and C atoms, respectively. Hydrogen atoms and solvents molecules were omitted for clarity. (bottom) XPS data for Zr(3d) and U(4f) regions for Zr6U0.87-Me2BPDC-8. 170x173mm (600 x 600 DPI)

ACS Paragon Plus Environment

Page 16 of 21

Page 17 of 21

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

Journal of the American Chemical Society

Figure 6. (top) Packing and metal nodes of Th6-Me2BPDC-10 and Th6U4-Me2BPDC-8. Red, purple, pink, and grey spheres represent Th, U, O, and C atoms, respectively. Hydrogen atoms and solvents molecules were omitted for clarity; (bottom) FTIR spectra of Th6-Me2BPDC-10 (red) and Th6U4-Me2BPDC-8 (purple). 178x190mm (600 x 600 DPI)

ACS Paragon Plus Environment

Journal of the American Chemical Society

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

Figure 7. Transmetallation in a molecular Zr-containing planar 15-membered macrocycle (top)62 and MOFs (bottom). 95x54mm (600 x 600 DPI)

ACS Paragon Plus Environment

Page 18 of 21

Page 19 of 21

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

Journal of the American Chemical Society

Figure 8. Packing and metal nodes of U6-Me2BPDC-8 and Th6-Me2BPDC-8. Insets show the color change occurred during cation exchange process. Red, purple, pink, and grey spheres represent Th, U, O, and C atoms, respectively. Hydrogen atoms and solvents molecules were omitted for clarity. 122x89mm (600 x 600 DPI)

ACS Paragon Plus Environment

Journal of the American Chemical Society

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

aThe red color indicates actinide location: red spheres represent An-based metal nodes; grey spheres – Zrbased metal nodes; grey solid sticks – organic linkers used for framework synthesis; blue springs – capping linker; red springs – capping linkers functionalized with an anchoring group; red icosahedra – non covalently bound actinide-containing guests (UO22+, Th4+). 228x346mm (600 x 600 DPI)

ACS Paragon Plus Environment

Page 20 of 21

Page 21 of 21

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

Journal of the American Chemical Society

Table of Content 94x53mm (600 x 600 DPI)

ACS Paragon Plus Environment