Supramolecular Assembly of Geometrically Unstable Hybrid Organic

Aug 2, 2019 - Please check again later. .... (32) U28 is a favored capsule topology when K is the only available .... Crystals of 1 or 2 were isolated...
0 downloads 0 Views 930KB Size
Subscriber access provided by GUILFORD COLLEGE

Article

Supramolecular Assembly of Geometrically Unstable Hybrid OrganicInorganic Uranyl Peroxide Cage Clusters and their Transformations Mengyu Xu, Hrafn Traustason, Fabrice Dal Bo, Sarah M. Hickam, Saehwa Chong, Lei Zhang, Allen G. Oliver, and Peter C Burns J. Am. Chem. Soc., Just Accepted Manuscript • DOI: 10.1021/jacs.9b05599 • Publication Date (Web): 24 Jul 2019 Downloaded from pubs.acs.org on July 25, 2019

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

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 27 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

Supramolecular Assembly of Geometrically Unstable Hybrid Organic-Inorganic Uranyl Peroxide Cage Clusters and their Transformations Mengyu Xu,a Hrafn Traustason,b Fabrice Dal Bo,a Sarah Hickam,a Saehwa Chong,a Lei Zhang,a Allen G. Oliver,b and Peter C. Burns* a,b a.

Department of Civil and Environmental Engineering and Earth Sciences, University of

Notre Dame, Notre Dame, Indiana, 46556, USA. b.

Department of Chemistry and Biochemistry, University of Notre Dame, Indiana, 46556,

USA ABSTRACT: An aromatic ligand was introduced into the synthesis of a uranyl peroxide polyoxometalate (POM) formulated as K32(UO2)19(O2)26(OH)2(C6H4P2O6)4⋅65 H2O that consists of a unique “open oyster” shaped structure (U19) with intramolecular H-bonds. In the solid state, K-π and π-π interactions, as well as K-O bonds enable the formation of a supramolecular network between U19 clusters. U19 adopts an incomplete fullerene topology and was utilized as a precursor from which the geometrically favored U24 structure was produced. A potassiumencapsulated U24 structure was obtained upon heating the solution containing U19. INTRODUCTION Polyoxometalates (POMs) are mostly anionic metal-oxo clusters of high-valence group V/VI cations.1-4 POMs range in size from sub‐nanoscale to protein scale and have structures and properties with applications in analytical and clinical chemistry, catalysis, medicine, and solidstate devices.1, 5-7 POMs exhibit substantial structural diversity owing to a variety of building blocks8 and many thousands of combinatorial possibilities.4 Design and control of the structure of POMs requires either bottom-up synthetic approaches or structural transformations from POM precursors. The synthetic variables include: 1) the concentration/type of metal oxide anion and heteroatom, 2) the presence of additional ligands, 3) the pH and ionic strength of the reaction mixture, 4) the presence of a reducing agent, and 5) the temperature of the reaction and processing.4 Generally, structural transformations of POM precursors are triggered by: 1) a change of pH caused by addition of acid/base,9 2) control of the pH-dependent equilibrium between POM monomers and dimers,10-11 3) modifying the structural “vacancy” sites of lacunary

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

POMs.12-16 The highly controllable structures and properties of POMs facilitate their use as building blocks for functional nanosystems and nanomachines at a larger scale (>10 nm).4 Controlling interactions between POM clusters during their assembly into extended structures is necessary and difficult when POMs are bridged by covalent bonds.5 To control supramolecular selfassembly, non-covalent forces such as Van der Waals interactions, hydrophobic binding, Hbonds, and/or metal-ligand interactions may be used.17 Functionalizing POMs by addition of organic ligands allows the use of H-bonds,18 cation-π interactions,19 and π-π interactions20 between clusters to guide the assembly of POMs into extended structures. High-valent actinides (An: actinide) often form linear (AnO2)+,2+ actinyl ions with multiplybound oxo ligands and uranyl- and neptunyl-based peroxo clusters have emerged as an extensive family of POMs.21-23 The majority of uranyl peroxide clusters are cages in which uranyl ions are bridged by peroxide or hydroxide ligands, with the uranyl ions arranged in topological squares, pentagons and hexagons.22 For example, U20 and U28 both have fullerene topologies, with the topology of U28 consisting of six hexagons in addition to 12 pentagons (Figure 1a and b, Un represents a cluster containing n uranyl polyhedra).24 U24 adopts the sodalite topology consisting of topological squares and hexagons (Figure 1c).25-28 U32 has a topology that contains squares, pentagons and hexagons (Figure 1d).25

Figure 1. Graphical representation of the topologies of (a) U20, (b) U28, (c) U24 and (d) U32. Linkages between uranyl ions are represented as yellow lines.

2

ACS Paragon Plus Environment

Page 2 of 27

Page 3 of 27 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

Uranyl peroxide clusters commonly form in alkaline aqueous solutions containing UO22+, H2O2 and alkali cations under ambient conditions.29 Varying the choice of alkali cation, UO22+/alkali cation ratio, pH, reaction time, and adding ligands during the synthetic process yields widely varied topologies and sizes of uranyl peroxide clusters.30-32 The topologies of uranyl peroxide cages synthesized using LiOH are controlled by reaction time and UO22+/Li ratio, forming U28 or U24.30 Controlling the UO22+/Na ratio promotes multiple cluster topologies in the NaOH system, yielding U20, U24, U28, and U32.32 U28 is a favored capsule topology when K is the only available counter cation, perhaps because K+ has an affinity with topological pentagons and hexagons formed by uranyl ions at room temperature during the self-assembly process.33-34 A structural transformation methodology from a uranyl peroxide cluster precursor attracts our attention as a possible approach to achieve structural variety in the KOH system. Here we report synthesis of (UO2)19(O2)26(OH)2(C6H4P2O6)432- (U19) and demonstrate a direct heating method that induces the structural transformation from U19 into the first K+-encapsulated U24 structure. U19 is an organic-functionalized lacunary cluster that has an incomplete U20 fullerene topology with intramolecular H-bonds and a unique “open oyster” shape. The presence of the aromatic ligand 1,2-bis(dimethoxyphosphoryl)benzene in U19 provides for non-covalent cation-π and π-π interactions between uranyl peroxide clusters, forming a higher-dimensional network of clusters. EXPERIMENTAL SECTION Caution! Experiments described herein were conducted using radioactive isotopically depleted uranium (238U, α = 4.467 MeV) by appropriately trained personnel in laboratories equipped and licensed to work with radioactive materials. General Considerations. Uranyl nitrate hexahydrate (UO2(NO3)2·6H2O, International BioAnalytical Industries, Inc.), hydrogen peroxide (30% aqueous solution, EMD Millipore), potassium hydroxide (KOH, Sigma- Aldrich), 1,2-bis(dimethoxyphosphoryl)benzene (C10H16O6P2, Alpha Aesar), and deuterium oxide (D2O, Cambridge Isotope Laboratories, Inc.) were obtained from commercial suppliers and used as received. Distilled and Millipore filtered water with a resistance of 18.2 MΩ⋅cm was used in all reactions. Isotopic labeling of the uranyl ion (U16O22+ → U18O22+) was conducted by a UV light-induced oxygen exchange reaction using 97% H218O (Cambridge Isotope Laboratories, Inc.) and was confirmed via solution-mode Raman

3

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

spectroscopy (Supporting Information).35 C6H4PO3H2 (BzDPA). Ten g of 1,2-bis(dimethoxyphosphoryl)benzene was dissolved in 20 mL of 12 M HCl and refluxed for 8 h. The product was washed with cold 2 M HCl, filtered, and dried at 60 °C for 12 hours.36 White crystals of benzene-1,2-diphosphonic acid (BzDPA) were formed in high yield (90%). 1H NMR (400 MHz, D2O): δ = 7.85, 2H, m; 7.52, 2H, m. 13C NMR (100 MHz, D2O): δ = 131.40, 2 CH, vt, Japp = 5; 132.86, 2 CH, vt, Japp = 12; 133.43, 2 CP, dd, 1J CP

= 175, 2JCP = 10. 31P NMR (162 MHz, D2O): δ = 14.32, 2 P, s.

K32(UO2)19(O2)26(OH)2(C6H4P2O6)4⋅65H2O (1). 0.1 mL of an aqueous 0.5 M UO2(NO3)2 solution and 0.1 mL of 30% H2O2 solution were combined in a 5 mL glass vial, resulting in precipitation of studtite, [UO2(O2)(H2O)2](H2O)2. Subsequently, 0.2 mL of 4.0 M aqueous KOH solution was added to dissolve the precipitate. Then 0.15 mL 0.2 M BzDPA solution was added and 4.0 M KOH solution was titrated to obtain pH 9.5. The resulting reaction solution was left open to the air and orange block-shaped crystals of 1 formed after four hours. Yield of 1: 30% on the basis of uranium. K32(U18O2)19(O2)26(OH)2(C6H4P2O6)4⋅65H2O (1*). 18O isotopically enriched 1* was obtained following the same method as for 1, except using 0.5 M U18O2(NO3)2. K24(UO2)24(O2)24(OH)24⋅nH2O (2). Twenty mg of 1 was dissolved in 0.5 mL 18MΩ H2O and the resulting solution was heated in a 23 mL capacity Teflon-lined stainless-steel Parr reaction vessel at 80°C for eight hours before cooling to room temperature. Yellow diamond-shaped crystals of 2 formed from the resulting solution after two days. Yield of 2: 10%. Elemental Analysis. Ten mg of 1 or 2 were dissolved in 0.5 mL 18MΩ H2O and the resulting solution was diluted in a 5% HNO3 matrix with yttrium added as an internal standard to monitor for instrument drift. Measurements were done using a Perkin Elmer Optima 8000 inductively coupled plasma optical emission spectrometer (ICP-OES) and the concentration of each element was calculated from calibration curves of standards with element concentrations in the range of 0.1 to 50 ppm. The reported value with uncertainty for each element is derived from the mean value and standard deviation of more than nine replicate measurements. The U: P: K ratios of 1 are 19: 7.90 ± 0.41: 32.11 ± 0.56, and the U: K ratios of 2 are 24: 20.38 ± 0.59. Crystallographic Studies. Crystals of 1 or 2 were isolated from their mother solutions and placed in oil to prevent their dehydration. A single crystal of each was selected using a polarized light microscope and was placed on a cryoloop for single crystal X-ray diffraction studies. 4

ACS Paragon Plus Environment

Page 4 of 27

Page 5 of 27 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

Diffraction data were measured using a Bruker APEX II Quazar diffractometer equipped with a micro source sealed tube with multilayer optics that produced a monochromated MoKα X-ray beam. Data were collected using frame widths of 0.5° in ω and an exposure time of 10 s for each frame. Data integration and corrections for Lorentz, polarization and background effects were performed using the Bruker APEX3 software package37 and corrections for absorption were applied using SADABS.38 The initial structure solutions were obtained by direct methods and least-squares refinements were performed by SHELXTL.39 The structure of 1 contains 19 symmetrically unique U sites, 18 of which are fully occupied by U. The site occupancy of U19 initially refined to 0.683(7), and U19 was subsequently replaced by two closely-spaced cation sites occupied by U (U19) and K (K19A). Refinement of the site occupancies of U19 and K19A with their total occupancy constrained to 1.0 yielded 0.608(9) U and 0.392(9) K and a U19-K19A separation of 0.33 Å. The interstitial regions of the structure of 1 contain somewhat disordered K and H2O that were tentatively assigned as partially occupied sites. The phenyl rings of 1 composed of C1-C24 were constrained to the vertices of rigid hexagons with C-C distances held at 1.39 Å. Hydrogen atoms were included on these phenyl rings as riding atoms with the constraint that C-H be 0.93 Å and Uiso(H) is 1.2 times larger than Ueq(C). Selected crystallographic information of 1 and 2 are listed in Table S2 and S6, respectively. Atomic coordinates and additional crystallographic information are provided in the Supporting Information (CIF). Bond valences incident upon cation and anion sites were calculated using literature parameters.40-41 Raman Spectroscopy. Raman spectra were measured using a Bruker Sentinel system with fiber optics and a video-assisted Raman probe equipped with a 785 nm, 148 mW laser and a high-sensitivity, TE-cooled CCD detector. Spectra were collected over the range 80-3200 cm-1. Crystalline U19 was placed under a Nikon optical microscope attached to the Raman probe for data collection. Spectral deconvolution of the 650-900 cm-1 region of 1 was done by fitting seven sets of Gaussian functions. Spectra were also collected using this instrument for solutions contained within glass vials. In cases where solutions initially contained a precipitate, they were first centrifuged before collecting Raman spectra. Heat Treatment of Aqueous U19. Eight solutions were prepared for study that each had 20 mg of 1 dissolved in 0.5 mL 18MΩ H2O. One of these solutions was left sitting at room

5

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

temperature for 8 hours as a control, and the others were heated at different temperatures ranging from 40°C to 160°C for 8 hours with ramping and cooling rates of 1°C/min. Yellow precipitates formed in the solutions that were heated above 80°C and were analyzed using powder X-ray diffraction. Precipitates were recovered by centrifugation and the resulting clear solutions were analyzed by ICP-OES to determine the concentration of uranium. 31P nuclear magnetic resonance (NMR) spectra, electrospray ionization mass spectra (ESI-MS) and Raman spectra were collected for each solution following heating. The reaction solutions were then left open to the air under ambient conditions to allow slow evaporation of the solvent. Crystals of 2 were recovered from the solutions heated at 80°C and 100°C after they were left standing in air for two days. 31P

Nuclear Magnetic Resonance Spectroscopy (31P NMR). 31P NMR measurements were

performed for aqueous solutions using a 600 MHz Varian INOVA spectrometer (11.74 T) with a pulse length of 14.8 ms, 128 scans, and a relaxation time (d1) of 10 s. Diffusion-ordered spectroscopy (DOSY) data were collected using the modified bipolar pulse stimulated echo (BPPSTE) sequence at 22°C, which were implemented in the Varian VnmrJ® version 2.2 revision C software package (AgilentTechnologies, Santa Clara, CA, USA).42 The gradients were increased from 2.3 to 50.7 G ∙ cm-1 in 10 steps using 256 scans. The gradient pulse (δ) was set to 3.0 ms and the diffusion time (Δ) was set to 0.1 s. Diffusion coefficients for resolved 31P signals were extracted from decay curves using the ‘peak height fit’ in the DOSY Transform module of MestReNova 11.0.4. The reported diffusion coefficients are the mean values with standard deviations of five repetitions. All samples were measured with a coaxial insert containing D2O and 0.1 M H3PO4. 31P chemical shifts were referenced to H3PO4 at 0 ppm. Electrospray Ionization Mass Spectrometry (ESI-MS). ESI-MS spectra were acquired using a Bruker qTOF time-of-flight spectrometer. An aqueous solution of U19 was prepared at a 10-20 μM concentration and was directly infused into the ionization source using a syringe pump at a flow rate of 4 μL/min. TOF data were acquired in negative-ion mode at a capillary voltage of 3600 V, end plate offset of -500 V, 0.8 bar dry gas, 1.2 L/min desolvation gas, and 180°C dry gas temperature. Additional Characterization Data. Details of thermogravimetric analysis (TGA), infrared spectroscopy (IR), and powder X-ray diffraction (P XRD) methods and results are available in Supporting Information.

6

ACS Paragon Plus Environment

Page 6 of 27

Page 7 of 27 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

RESULTS AND DISCUSSION Cluster synthesis and cluster salt composition of compound 1. Combining an aqueous solution of uranyl nitrate, hydrogen peroxide, BzDPA and potassium hydroxide at pH ~9.5 under ambient conditions resulted in the crystallization of 1 within four hours. Crystallization of 1 occurs more quickly than for most salts of uranyl peroxide clusters, which often only crystallize after significant solvent evaporation and several days. X-ray diffraction data provided the crystal structure of 1 that contains well-ordered positions of U, O, P and C atoms arranged in cage clusters containing 19 UO22+ uranyl ions (Figure 2). According to the crystal structure determination, the anionic cage of 1 has composition [(UO2)19(O2)26(O)2(C6H4P2O6)4]34-, where O2 corresponds to peroxide. The structure determination also revealed locations of 27 K sites distributed inside and outside of the cluster cage. Of these, 10 have refined occupancies less than unity, and may contain a mixture of K and H2O. Despite disorder of the H2O positions, 59 H2O sites were located, with 49 of these having occupancies modelled as less than unity. The elemental analysis of 1 yielded a U: P: K ratio of 19: 7.90 ± 0.41: 32.11 ± 0.56, indicating ~32 K cations per formula unit. Thermogravimetric analysis (TGA) yielded ~65 H2O per formula unit in 1 (Supporting Information). Thus, the formula of 1 is assigned as K32(UO2)19(O2)26(OH)2(C6H4P2O6)4(H2O)65, with the occurrence of hydroxyl discussed below. Structure of U19. The cage cluster in 1 consists of 19 uranyl hexagonal bipyramids and four BzDPA ligands (Figure 2a and b, designated U19). In the discussion that follows the cluster is described for the case where the U19 polyhedron is occupied by U, although the structure determination indicates that in about 40% of the clusters of the studied crystal this polyhedron is occupied by K. Each U(VI) cation is strongly bonded to two oxygen atoms in a trans arrangement, forming the usual UO22+ uranyl ion with an average uranium to yl-oxygen (U-Oyl) bond length of 1.80 Å. Each uranyl ion is coordinated to six oxygen-containing ligands arranged at equatorial positions of hexagonal bipyramids, and the average uranium to equatorial-oxygen (U-Oeq) bond length is 2.36 Å. Nine uranyl polyhedra contain two peroxo ligands that are bidentate to the uranyl ion in a cis arrangement and that correspond to two of the equatorial edges of hexagonal bipyramids. Eight of these uranyl ions are also coordinated by two oxygen atoms of diphosphonic groups, and the atom U19 is coordinated by two O2- ligands that are likely protonated (see below). The other uranyl ions have three bidentate peroxo ligands that are

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

mostly shared with adjacent uranyl polyhedra. However, in the cases of the U11, U15, U16 and U17 polyhedra, one of the peroxo ligands is non-bridging. Eight phosphorus cations are tetrahedrally coordinated by oxygen atoms and carbon atoms that are part of four different benzene cycles. U19 is the first uranyl cluster containing 19 uranyl cations and has two distinct building units that are linked by four BzDPA groups. One of these has eight uranyl polyhedra assembled into two five-membered rings, and one has 11 uranyl polyhedra assembled into three five-membered rings (Figure 2c and d). The linkages between uranyl polyhedra and BzDPA groups form distorted five-membered rings, giving a novel overall 19:57 topology (Figure 2c, d and e).

Figure 2. Polyhedral representations of the U19 cage cluster in 1 from (a) front and (b) back. (c) Graphical representation of the topology of U19, of which the building blocks are five-membered rings. Combined polyhedral ball-and-stick representations of (d) the five-membered ring of uranyl ions bridged by peroxide ligands and (e) the five-membered ring of uranyl ions bridged by BzDPA and peroxide ligands. In (a) and (b), uranyl polyhedra are yellow, phosphorus polyhedra are magenta, and carbon-carbon bonds are black. In (b), the yellow lines correspond to linkages between uranyl ions through peroxide ligands, the magenta lines designate linkages between uranyl ions through BzDPA ligands. In (d) and (e), uranium and oxygen atoms are shown as yellow and red spheres, respectively.

Intramolecular H-bonds. Four inorganic lacunary uranyl peroxide clusters have been reported to date: the open-bowl-shaped U1643 and crown-shaped U20R,43 U24R,43 and U32R.44 U19 is an organic-functionalized lacunary cluster that has an incomplete U20 fullerene topology with two open sides (Figure 3a). Compared to U20, omission of one uranyl ion results in an open side

8

ACS Paragon Plus Environment

Page 8 of 27

Page 9 of 27 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

on U19 where it is terminated by the triperoxide uranyl hexagonal bipyramids U11, U15, and U17, each of which includes a non-bridging peroxide ligand. These non-bridging peroxide ligands are bonded to potassium cations and also likely accept H-bonds donated by interstitial water to satisfy their bond-valence deficiency of ~0.65 valence units (vu) calculated in the absence of H bonds. Previously reported crown-shaped uranyl peroxide clusters also feature nonbridging peroxide groups that bond to interstitial cations and accept H bonds from interstitial water.43-44 On the opposite side of U19 (Figure 3c), the U16 polyhedron is terminated by peroxide oxygen atoms O97 and O85 and this peroxide is not shared by neighboring U19, whereas its equivalent is in the case in the U20 cluster structure, leaving a gap between these two uranyl polyhedra and producing an ‘open-oyster’ shape. The distance between U16 and U19 is ~1.5 Å longer than the typical U-U distance of uranyl ions bridged by peroxide ligands. Figure 3. (a) U19 has an incomplete U20 fullerene topology. Details of the two open sides of U19: (b) Three non-bridging

peroxide ligands truncate the cluster and likely accept H bonds from interstitial H2O. (c) U19 contains two non-bridging oxygen atoms O110 and O117 that donate H bonds to the accepters O97 and O85. Legend as in Figure 2. Potassium atoms are shown as purple spheres.

O110 and O17, which are not peroxide oxygen atoms, terminate the U19 diperoxide hexagonal bipyramid, and each is also bonded to one potassium cation. The calculated bondvalence sums at these O sites from the bonds to the uranium and potassium cations are 0.52 vu for O110 and 0.63 vu for O17. This situation is consistent with O110 and O17 being occupied by

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

either H2O or OH-, as they may also accept one or two H bonds that would be important if they are occupied by OH-. The O110-O97 and O17-O85 distances across the opening in U19 (Figure 3c) are 2.71 Å and 2.61 Å, consistent with intramolecular H bonds, and the bond-valence strengths of the U-O bonds of O97 and O85 are 0.62 vu and 0.61 vu, consistent with them acting as H bond acceptors. Based on the elemental analysis and the requirement of electroneutrality, a proton is assigned to each of O110 and O17. Intermolecular Interaction Between U19 Clusters. In compound 1, there are T-shaped π-π stacking interactions, which are also identified as C-H···π interactions, between BzDPA ligands of adjacent U19 clusters that are characterized by centroid-centroid separations of ~5.11 Å and ~97° between the normals to the ring plane (Figure 4a).45 A computational study suggested that T-shaped and slipped-parallel aromatic dimers are both enthalpically favored, and the T-shaped dimers have the largest interaction energy when the intermolecular distance is 5.0 Å.46 The cluster dimers are connected to other clusters with different orientations through K-O bonds and a K-π interactions (Figure 4a). The C-H···π interactions facilitate the formation of chains of cage clusters in the (010) plane and linked by K-O bonds (Figure 4b). K-π and K-O interactions link the chains in the (101) plane and form the overall supramolecular network of 1 (Figure 4b). We propose that the supramolecular interactions between clusters described here facilitate the rapid crystallization of 1, which at four hours is much faster than is typical for salts of uranyl peroxide cage clusters. Raman Spectroscopy. The Raman spectra of 1 and its O18-enriched variety and their peak assignments are shown in Figure 5. Peroxide ligands bonded to uranyl produce Raman bands in the range of 840-870 cm-1,47 and the stretch of η2-peroxide appears at 861 cm-1 in the Raman spectrum of studtite, [UO2(O2)(H2O)2](H2O)2.48 In Raman spectra of salts of clusters Li-U24 and Li/K-U60, which contain uranyl diperoxide hexagonal bipyramids only, peaks near 850 cm-1 were assigned to the O-O symmetric stretch (νs, O-O),49 although a different study assigned this vibrational mode to uranyl-bridging hydroxyl groups.50 Here, we assigned the mode at 861 cm-1 to the νs, O-O from the eight diperoxide uranyl units since there are no uranyl-bridging hydroxyl groups in this structure. The mode at 825 cm-1 is assigned to the νs, O-O of the ten triperoxide uranyl units, consistent with the reported value of νs, O-O in uranyl peroxide clusters composed

10

ACS Paragon Plus Environment

Page 10 of 27

Page 11 of 27 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

of uranyl triperoxide hexagonal bipyramids that occur in the range of 800-850 cm-1.30, 51 The mode at 801 cm-1 is assigned to the symmetric stretch of uranyl ions (νs, O-U-O), in line with the reported Raman spectra of uranyl peroxide clusters.30, 49 The mode at 757 cm-1 is tentatively assigned to the symmetric vibration of uranyl (νs, O-U-O) from uranyl ions coordinated to nonbridging peroxide groups, U11, U15 and U17, based on the corresponding values in the spectra of uranyl triperoxide monomers that occur in the range of 680 to 740 cm-1. The mode at 680 cm-1 is attributed to the symmetric stretching vibration of PO3 (νs, P-O-P).

Figure 4. (a) C-H···π (black dashed lines) and K-π (purple dashed lines) interactions between adjacent U19 clusters; space-filling diagram of these interactions. (b) Graphic representation of the network composed of cluster chains. The U19 clusters are represented as yellow spheres and are connected through K-O bonds (grey lines).

To support the Raman shift assignments, the spectrum of O18-enriched crystalline salt of 1 (1*) was collected. Note that the O18 label was placed on the uranyl ion, and no exchange of this label into peroxide is expected. The lack of response to isotopic substitution of bands at 825 and 861 cm-1 is consistent with their assignment to peroxide. The growth of the mode at 757 cm-1 and the decrease of that at 801 cm-1 upon isotopic substitution indicates that these are from an isotopic pair of uranyl ions, represented as (O-U-O)A here. Two modes appear in the region of

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

780-820 cm-1 and are fitted at 786 and 801 cm-1 after the shifting of the (O-U-O)A peak. They are assigned to the vibration of O-O because they did not respond to the isotopic substitution at the uranyl oxygen sites. These two modes have a lower wavenumber as compared to the other peroxide peaks but are close to the values reported for the uranyl triperoxide monomers at around 810 cm-1;51 as such, these peaks may be due to non-bridging peroxide groups in 1. The mode at 718 cm-1 in the spectrum of 1* corresponds to the mode at 757 cm-1 in the spectrum of 1. This peak shift is in good agreement with our assignment above and this peak is designated as the (O-U-O)B peak. The empirical formulation RUO = 106.5ν1-2/3 + 0.575 developed by Bartlett52 correlates the experimentally observed wavenumber of νs, O-U-O (ν1) with the crystallographic U-Oyl bond lengths. The predicted value in this case is 801 cm-1, which is consistent with the experimental value of the (O-U-O)A peak. Polyhedron U19 has a long U-Oyl bond length of 1.86 Å and the empirically predicted peak position is 755 cm-1, which is in accord with the experimental value of the (O-U-O)B peak. The substitution of equatorially-bound water molecules by hydroxyl groups is expected to lead to weakening of uranyl bonds and is reflected in the redshift of Raman peaks.53

Figure 5. Raman spectra of isotopically neutral (black) and 18O-enriched (blue) crystals of 1.

Transformation from U19 to [(UO2)24(O2)24(OH)24]24- (U24). Inclusion of KOH in the aqueous uranyl peroxide system commonly yields U28 (K-U28), which earlier was attributed to a cation templating effect with the uranyl five-membered ring of fullerene-topology U28 favoring

12

ACS Paragon Plus Environment

Page 12 of 27

Page 13 of 27 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

coordination to K+.21, 33 Synthesis of uranyl peroxide clusters via structural transformation may provide opportunities to circumvent templating effects of counter cations during the formation of clusters. Addition of acid/base to trigger structural transformations is common in cluster chemistry9 but is not entirely suitable for uranyl peroxide clusters because these clusters often decompose upon significant change of pH. Recently it was observed that U60 in aqueous solution converts into U24 upon heating,54 prompting us to heat solutions of U19 in an attempt to trigger a transformation. Efforts to prepare other K-containing cluster structures from K-U28 through structural transformations triggered by addition of acid or base or heating thus far have failed and K-U28 persists in the solution. U19 is a potential POM precursor because: 1) it contains both uranyl diperoxide units and triperoxide units that are basic building units for most uranyl peroxide clusters,22 and 2) the defected U20 fullerene structure of U19 may be susceptible to structural transform into a geometrically more stable structure. Crystals of 2, which contain the well-known U24 cluster and K counter cations, were isolated from aqueous solutions of U19 after heat treatment from 60-100°C. U24 was first synthesized with Li+ as the counter cation and has since been reported containing Na+, Bi3+, Pb2+ and alkaline earth elements (Ca, Sr, Ba),26-28 as well as combinations of Li/Na and Li/K.55 U24 consists of 24 uranyl hexagonal bipyramids, each of which contains two peroxide ligands and two hydroxyl groups (Figure 6a).21 Single-crystal X-ray diffraction data indicate U24 in 2 has composition [(UO2)24(O2)24(OH)24]24-, where O2 corresponds to peroxide. Twenty-four K sites were located in the crystal structure that balance the charge of the cage cluster, although six symmetrically equivalent K sites are each split into four sites that are disordered with water oxygen atoms. Eight K cations are inside the U24 cage, where they are positioned inside the topological hexagons that are each defined by six uranyl ions, and these are arranged at the vertices of a cube (Figure 6b). Each of these encapsulated K+ cations are bonded to uranyl oxygen atoms and water (Figure 6c). A similar arrangement of the encapsulated counter ions was reported earlier for Pb8O64+ polyoxocations in U24, where Pb2+ is also located under the hexagonal windows,27 as well as the K cations in a salt of U24 that contains both Li and K.55

13

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 6. (a) Polyhedral representation of U24 contained in 2. (b) K+ cations in a cubic arrangement encapsulated in the cage of U24. (c) Each K+ is 8-coordinated by six uranyl oxygen atoms and two water oxygen atoms. Legend as in Figure 1 except water oxygen atoms are shown in cyan.

31P

NMR spectra were collected for room temperature solutions of U19 that had earlier been

heated in sealed reactor vessels to soak temperatures from 40 to 160 °C for 8 hours, as well as a control aged at room temperature for 8 hours (Figure 7a). The disassembly of U19 started in solutions heated to 40°C, as indicated by the decrease of the peak intensity of BzDPA ligands contained in U19 (signal A, 17.92 ppm) and increase of the peak intensity of disassociated BzDPA ligands (signal B, 12.20 ppm) in the solution. A phosphorus species is evident in the spectrum from the solution heated to 80°C as peak C (20.56 ppm) and corresponds to ~14% of the total P in solution. Signal A is absent in spectra collected for solutions heated above 100°C, indicating full disassembly of U19. The highest yield of crystals of 2 was obtained from the solution heated at 80°C (~10%). Time-resolved 31P NMR spectra were collected from a solution of U19 heated at 80°C for eight hours in the NMR spectrometer (Figure 7b). Signal C appeared within one hour of the onset of heating and intensified along with signal B over time, accompanied with a decrease and broadening of signal A. Diffusion-ordered spectroscopy (DOSY) measurements were done for the room-temperature solution that had been heated in a sealed vessel at 80°C for eight hours. From DOSY spectra, the measured diffusion coefficient (D) for signal A is (1.44 ± 0.46) × 10−10 m2 ∙ s−1, and that for signal C is (1.11 ± 0.17) × 10−10 m2 ∙ s−1 (Figure S3). The similarity of these diffusion coefficients indicates that the species that produces signal C is similar to U19 (which produces signal A) in size.

14

ACS Paragon Plus Environment

Page 14 of 27

Page 15 of 27 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

31

Figure 7. (a) P NMR spectra of aqueous solutions that initially contained U19 collected at room temperature after solution aging at room temperature or after heating of solutions in a sealed vessel for 8 hours at 40 to 160 °C and cooling to room temperature. (b) Time-resolved 31P NMR spectra collected for a solution that initially contained U19 that was heated at 80°C for 8 hours.

Crystals of K-U24 were isolated in highest yield from a solution that initially contained U19 after heating at 80°C for 8 hours (hereafter designated the reaction solution). Crystals of U19 and U24 were dissolved in water and ESI-MS spectra were collected for comparison with that of the reaction solution (Figure 8a). Aqueous solutions prepared by dissolving the salts of U19 and U24 each contain signals corresponding to species with charges of -6, -5 and -4. The spectrum collected for the reaction solution contains signals corresponding to charges of -7, -6 and -5. Peaks at m/z values 1336.3, 1641.3 and 2095.1 in the spectrum of solution into which U24 was dissolved are assigned to [K9H9(UO2)24(O2)24(OH)24]6-, [K14H5(UO2)24(O2)24(OH)24]5- and [K18H2(UO2)24(O2)24(OH)24(H2O)]4-, respectively. In the spectrum of U19, each peak represents the cluster ion (UO2)19(O2)26(OH)2(C6H4P2O6)432- with varying combinations of K+ and H+

15

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

counterions and water. For example, prominent peaks observed at m/z values 1261.2, 1529.0 and 2015.9 correspond to [K16H10(UO2)19(O2)26(OH)2(C6H4P2O6)4]6-, [K18H9(UO2)19(O2)26(OH)2(C6H4P2O6)4]5- and [K28(UO2)19(O2)26(OH)2(C6H4P2O6)4(H2O)2]4-, respectively. The spectrum of the solution prepared by dissolution of U19 salt followed by heating to 80°C (the reaction solution) indicates the presence of multiple species and includes weak signals that are attributable to U24. However, peaks observed at m/z values 1248.0, 1462.6, 1516.4 and 1781.9 are inconsistent with either U24 or U19, and likely correspond to uranyl peroxide clusters with or without organic ligands, of which the molecular mass ranges from 8700 to 9100 (see Supporting information, Table S1 for the assignment of all peaks in Figure 8a). Figure 8. (a) ESI-MS spectra of the solution into which crystals of U19 or U24 was dissolved and the reaction

solution from which crystals of U24 were isolated. (b) ESI-MS spectra of aqueous solutions that initially contained U19 collected at room temperature after solution aging at room temperature or after heating of solutions in sealed vessels for 8 hours at 40 to 160 °C and cooling to room temperature.

16

ACS Paragon Plus Environment

Page 16 of 27

Page 17 of 27 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

As for the 31P NMR experiment, ESI-MS spectra were also collected for room temperature solutions of U19 that had earlier been heated in sealed vessels from 40 to 160 °C for 8 hours, as well as a solution aged at room temperature for 8 hours (Figure 8b). The spectra indicate that peaks attributed to U19 persist in solutions heated up to 60 C, although peaks associated with unknown clusters are dominant for solutions heated to 40 C or higher. Raman spectra were collected for the various heated solutions that were initially prepared by dissolving the salt of U19 (Figure 9). The salt of U19 (1) produces a Raman mode at 825 cm-1 assigned to the νs, O-O of the uranyl triperoxide units, and this is also present in the spectrum of the solution produced by dissolving the salt of U19. The Raman spectra of an aqueous solution of U24 and the reaction solution (heated to 80 C) contain a mode at 848 cm-1 that is assign to the stretch of peroxide in the uranyl diperoxide units. The aqueous solution of U24 gives a uranyl vibration mode at 812 cm-1, whereas the reaction solution and a solution into which the salt of U19 was dissolved in water produce uranyl vibration modes at 804 cm-1. The signals at 1040 cm-1 are assigned to the asymmetric stretching of PO3 (νas, P-O-P). The mode at 1067 cm-1 for the reaction solution corresponds to the asymmetric stretch of PO3 (νas, P-O-P) of dissociated BzDPA ligands. In the room temperature spectrum of the solution heated to 60°C for 8 hours, the signal at 825 cm-1 is reduced, while the signals at 848 cm-1 and 1067 cm-1 are strengthened relative to the spectrum of an unheated solution, indicating conversion from uranyl triperoxide species to uranyl diperoxide species and the dissociation of BzDPA ligands. Increasing temperature fosters formation of uranyl diperoxide species until the temperature is above 100°C. The modes attributable to uranyl peroxide cluster species decrease above 80°C, and are no longer detectable for solutions heated at 140°C for 8 hours.

17

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 9. (a) Raman spectra of solutions into which crystals of U19 or U24 were dissolved and the reaction solution from which crystals of U24 were isolated. (b) Raman spectra of aqueous solutions that initially contained U19 collected at room temperature after solution aging at room temperature or after heating of solutions in sealed vessels for 8 hours at temperatures from 40 to 160 °C and cooling to room temperature.

U19 is composed of topological pentagons of uranyl di or tri-peroxide units, whereas U24 contains topological squares and hexagons of uranyl diperoxide dihydroxide units. Previous studies suggested assembly of uranyl triperoxide monomers into uranyl peroxide capsules through tetramer, pentamer, and hexamer intermediates.50, 56 The significant topological differences between U19 and U24 require nearly complete disassembly of U19 into uranyl peroxide monomers or dimers followed by a reassembly into U24. When the aqueous solution of U19 is heated between 60°C and 100°C, a disassembly of U19 is induced. Raman spectra suggest that uranyl diperoxide species dominate in the solution and no significant Raman signal corresponding to uranyl triperoxide species was observed. Conversion of uranyl triperoxide monomers to diperoxide monomers can occur through the disproportionation of peroxide ligands 18

ACS Paragon Plus Environment

Page 18 of 27

Page 19 of 27 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

to form two bridging hydroxyl ligands and free molecular oxygen.31 Hence, a disproportionation reaction may occur during heating: 2 UO2(O2)34- + 2 H2O → 2 UO2(O2)2(OH)24- + O2↑.31 31P NMR spectra suggest the presence of a phosphorous-containing cluster in the reaction solution where the highest yield of U24 crystals was obtained. ESI-MS indicate that a cluster of mass 8.7 – 9.1 KDa is present in the reaction solution. This unidentified cluster likely forms due to assembly of uranyl diperoxide units and BzDPA ligands according to the 31P NMR and ESI-MS results. A precipitate forms when an aqueous solution that initially contained U19 is heated above 100°C (see Supporting Information, Figure S4 for the ICP-OES results of the concentration of uranium in solutions after heat treatment). Powder X-ray diffraction patterns collected for precipitates formed in solutions heated from 140 to 200°C indicated the major phase is metaschoepite ((UO2)4O(OH)6)(H2O)5, whereas precipitates formed at lower temperatures are poorly crystalline (see Supporting Information, Figure S5 for the P XRD results).

CONCLUSIONS The self-assembly of uranyl peroxide polyhedra with benzene-1,2-diphosphonic acid formed U19 that features a defected fullerene topology with 19 uranyl ions. This is the first example of a uranyl peroxide cluster that contains terminal hydroxyl ligands and intramolecular H-bonds. With the presence of benzene-containing ligands, K-π and π-π interactions along with K-O bonds connect U19 clusters in the solid state to form a POM-based supramolecular network. These interactions appear to accelerate the rate of cluster crystallization relative to most uranyl peroxide cage clusters. The functionalizing of uranyl peroxide clusters using organic ligands as shown here may provide a route to uranyl peroxide cluster-based metal organic frameworks (MOFs). U19 was selected to be a POM precursor to obtain other POMs via structural transformation. Direct heating of an aqueous solution was applied to trigger the structural transformation of U19 through a disassembly and reassembly process. In response to the increasing temperature, U19 that was dissolved in water fragmented and the uranyl polyhedra reassembled into U24 above 60°C. Crystallization of U24 from the heated solution provided the first example of encapsulation of only potassium and water in U24. Above 100°C, the uranyl peroxide cluster decomposes and metashoepite precipitates. U28, on the other hand, is inert when exposed to the same reaction

19

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

conditions, which indicates the reactivity of U19 may result from its unstable geometry. Deconstruction of a geometrically unstable POM may generate building blocks that are not commonly formed from normal synthesis approaches; in this case, uranyl diperoxide units. Demonstrated by the successful synthesis of U24 from U19, we propose a workflow here to obtain kinetically or thermodynamically unfavorable POM structures, which may be expanded to the transition metal POMs synthesis. Step 1: introduce additional organic or inorganic ligands into the POM structure to create geometrically unstable structures. Step 2: Induce fragmentation of the unstable structure into basic building blocks that assemble into a new structure. ASSOCIATED CONTENT Supporting Information The Supporting Information is available free of charge on the ACS Publications website at DOI:. Crystallographic information file for 1 and 2 (CIF) Thermogravimetric analysis (TGA), infrared spectroscopy (IR), Diffusion-ordered Spectroscopy (DOSY), Electrospray Ionization Mass Spectrometry (ESI-MS), Concentration of U in the resulting solution from the heating experiments, and additional crystallographic information (PDF) AUTHOR INFORMATION Corresponding Author Peter C. Burns, [email protected] Notes The authors declare no competing financial interests. ACKNOWLEDMENTS This work was funded by the Chemical Sciences, Geosciences and Biosciences Division, Office of Basic Energy Sciences, Office of Science, U.S. Department of Energy, Grant No. DE-FG02– 07ER15880. We thank Dr. Jaroslav Zajicek and Dr. Evgenii Kovrigin for assisting with NMR measurements, and Dr. William Boggess for helping with ESI-MS analyses. We also thank Dr. Seth N. Brown for his advice on NMR spectra analyses. The Magnetic Resonance Research 20

ACS Paragon Plus Environment

Page 20 of 27

Page 21 of 27 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

Center, Center for Environmental Science and Technology, Materials Characterization Facility, and Mass Spectrometry and Proteomics Facility at the University of Notre Dame provided instrumentation used in this work.

21

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

REFERENCES (1) Pope, M. T.; Müller, A., Polyoxometalate chemistry: an old field with new dimensions in several disciplines. Angew. Chem., Int. Ed. 1991, 30 (1), 34-48. (2) Hill, C. L., Introduction: polyoxometalates multicomponent molecular vehicles to probe fundamental issues and practical problems. Chem. Rev 1998, 98 (1), 1-2. (3) Long, D.-L.; Burkholder, E.; Cronin, L., Polyoxometalate clusters, nanostructures and materials: From self assembly to designer materials and devices. Chem. Soc. Rev. 2007, 36 (1), 105-121. (4) Long, D.-L.; Tsunashima, R.; Cronin, L., Polyoxometalates: Building Blocks for Functional Nanoscale Systems. Angew. Chem., Int. Ed. 2010, 49 (10), 1736-1758. (5) Long, D.-L.; Cronin, L., Towards Polyoxometalate-Integrated Nanosystems. Chem. Eur. J. 2006, 12 (14), 3698-3706. (6) Müller, A.; Beckmann, E.; Bögge, H.; Schmidtmann, M.; Dress, A., Inorganic Chemistry Goes Protein Size: A Mo368 Nano-Hedgehog Initiating Nanochemistry by Symmetry Breaking. Angew. Chem., Int. Ed. 2002, 41 (7), 1162-1167. (7) Rhule, J. T.; Hill, C. L.; Judd, D. A.; Schinazi, R. F., Polyoxometalates in Medicine. Chem. Rev. 1998, 98 (1), 327-358. (8) Yamase, T.; Pope, M. T., Polyoxometalate chemistry for nano-composite design. Springer Science & Business Media, 2006. (9) Liu, H.; Song, C. Y.; Huang, R. W.; Zhang, Y.; Xu, H.; Li, M. J.; Zang, S. Q.; Gao, G. G., Acid–Base‐Triggered Structural Transformation of a Polyoxometalate Core Inside a Dodecahedrane‐like Silver Thiolate Shell. Angew. Chem., Int. Ed. 2016, 55 (11), 3699-3703. (10) Kholdeeva, O. A.; Maksimov, G. M.; Maksimovskaya, R. I.; Kovaleva, L. A.; Fedotov, M. A.; Grigoriev, V. A.; Hill, C. L., A Dimeric Titanium-Containing Polyoxometalate. Synthesis, Characterization, and Catalysis of H2O2-Based Thioether Oxidation. Inorg. Chem. 2000, 39 (17), 3828-3837. (11) Rhule, J.; Weinstock, I.; Hill, C., Synthesis, X-ray structure, and hydrolytic chemistry of the highly potent antiviral polyniobotungstate A-α-[Si2Nb6W18O77]8–. Chem. Commun. 1999, (17), 1651-1652. (12) Kikukawa, Y.; Suzuki, K.; Yamaguchi, K.; Mizuno, N., Synthesis, Structure Characterization, and Reversible Transformation of a Cobalt Salt of a Dilacunary γ-Keggin 0

ACS Paragon Plus Environment

Page 22 of 27

Page 23 of 27 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

Silicotungstate and Sandwich-Type Di- and Tetracobalt-Containing Silicotungstate Dimers. Inorg. Chem. 2013, 52 (15), 8644-8652. (13) Maruyama, T.; Kikukawa, Y.; Sakiyama, H.; Katayama, M.; Inada, Y.; Hayashi, Y., A highly-flexible cyclic-decavanadate ligand for interconversion of dinuclear- and trinuclearcobalt(ii) and manganese(ii) cores. RSC Adv. 2017, 7 (60), 37666-37674. (14) Saku, Y.; Sakai, Y.; Nomiya, K., Relation among the 2:2-, 1:1- and 1:2-type complexes of hafnium(IV)/zirconium(IV) with mono-lacunary α2-Dawson polyoxometalate ligands: Synthesis and structure of the 2:2-type complexes [{α2-P2W17O61M(μ-OH)(H2O)}2]14− (M=Hf, Zr). Inorg. Chim. Acta 2010, 363 (5), 967-974. (15) Kuriyama, Y.; Kikukawa, Y.; Suzuki, K.; Yamaguchi, K.; Mizuno, N., Water- and Temperature-Triggered Reversible Structural Transformation of Tetranuclear Cobalt(II) Cores Sandwiched by Polyoxometalates. Chem. Eur. J. 2016, 22 (12), 3962-3966. (16) Suzuki, K.; Hanaya, T.; Sato, R.; Minato, T.; Yamaguchi, K.; Mizuno, N., Hexanuclear tin(ii) and mixed valence tin(ii,iv) oxide clusters within polyoxometalates. Chem. Commun. 2016, 52 (70), 10688-10691. (17) Philp, D.; Stoddart, J. F., Self‐assembly in natural and unnatural systems. Angew. Chem., Int. Ed. 1996, 35 (11), 1154-1196. (18) Pradeep, C. P.; Long, D.-L.; Newton, G. N.; Song, Y.-F.; Cronin, L., Supramolecular Metal Oxides: Programmed Hierarchical Assembly of a Protein-Sized 21 kDa [(C16H36N)19{H2NC(CH2O)3P2V3W15O59}4]5− Polyoxometalate Assembly. Angew. Chem., Int. Ed. 2008, 120 (23), 4460-4463. (19) Rosnes, M. H.; Musumeci, C.; Pradeep, C. P.; Mathieson, J. S.; Long, D.-L.; Song, Y.-F.; Pignataro, B.; Cogdell, R.; Cronin, L., Assembly of Modular Asymmetric Organic− Inorganic Polyoxometalate Hybrids into Anisotropic Nanostructures. J. Am. Chem. Soc. 2010, 132 (44), 15490-15492. (20) Song, Y. F.; Long, D. L.; Cronin, L., Noncovalently connected frameworks with nanoscale channels assembled from a tethered polyoxometalate–pyrene hybrid. Angew. Chem., Int. Ed. 2007, 46 (21), 3900-3904. (21) Burns, P. C.; Kubatko, K.-A.; Sigmon, G.; Fryer, B. J.; Gagnon, J. E.; Antonio, M. R.; Soderholm, L., Actinyl Peroxide Nanospheres. Angew. Chem., Int. Ed. 2005, 117 (14), 21732177.

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

(22) Qiu, J.; Burns, P. C., Clusters of actinides with oxide, peroxide, or hydroxide bridges. Chem. Rev. 2012, 113 (2), 1097-1120. (23) Burns, P. C.; Nyman, M., Captivation with encapsulation: a dozen years of exploring uranyl peroxide capsules. Dalton Trans. 2018, 47 (17), 5916-5927. (24) Sigmon, G. E.; Ling, J.; Unruh, D. K.; Moore-Shay, L.; Ward, M.; Weaver, B.; Burns, P. C., Uranyl−Peroxide Interactions Favor Nanocluster Self-Assembly. J. Am. Chem. Soc. 2009, 131 (46), 16648-16649. (25) Burns, P. C.; Kubatko, K. A.; Sigmon, G.; Fryer, B. J.; Gagnon, J. E.; Antonio, M. R.; Soderholm, L., Actinyl peroxide nanospheres. Angew. Chem., Int. Ed. 2005, 117 (14), 21732177. (26) Zanonato, P. L.; Di Bernardo, P.; Fischer, A.; Grenthe, I., Chemical equilibria in the UO22+– H2O2–F−/OH− systems and possible solution precursors for the formation of [Na6(OH2)8]@[UO2(O2)F]2418− and [Na6(OH2)8]@[UO2(O2)OH]2418− clusters. Dalton Trans. 2013, 42 (28), 10129-10137. (27) Renier, O.; Falaise, C.; Neal, H.; Kozma, K.; Nyman, M., Closing Uranyl Polyoxometalate Capsules with Bismuth and Lead Polyoxocations. Angew. Chem., Int. Ed. 2016, 128 (43), 1367813682. (28) Falaise, C.; Hickam, S. M.; Burns, P. C.; Nyman, M., From aqueous speciation to supramolecular assembly in alkaline earth-uranyl polyoxometalates. Chem. Commun. 2017, 53 (69), 9550-9553. (29) Nyman, M.; Burns, P. C., A comprehensive comparison of transition-metal and actinyl polyoxometalates. Chem. Soc. Rev. 2012, 41 (22), 7354-7367. (30) Falaise, C.; Nyman, M., The Key Role of U28 in the Aqueous Self-Assembly of Uranyl Peroxide Nanocages. Chem. Eur. J. 2016, 22 (41), 14678-14687. (31) Miró, P.; Vlaisavljevich, B.; Gil, A.; Burns, P. C.; Nyman, M.; Bo, C., Self‐Assembly of Uranyl–Peroxide Nanocapsules in Basic Peroxidic Environments. Chem. Eur. J. 2016, 22 (25), 8571-8578. (32) Hickam, S.; Aksenov, S. M.; Dembowski, M.; Perry, S. N.; Traustason, H.; Russell, M.; Burns, P. C., Complexity of Uranyl Peroxide Cluster Speciation from Alkali-Directed Oxidative Dissolution of Uranium Dioxide. Inorg. Chem. 2018, 57 (15), 9296-9305.

2

ACS Paragon Plus Environment

Page 24 of 27

Page 25 of 27 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

(33) Miró, P.; Pierrefixe, S.; Gicquel, M.; Gil, A.; Bo, C., On the Origin of the Cation Templated Self-Assembly of Uranyl-Peroxide Nanoclusters. J. Am. Chem. Soc. 2010, 132 (50), 1778717794. (34) Arteaga, A.; Zhang, L.; Hickam, S.; Dembowski, M.; Burns, P. C.; Nyman, M., Uranyl– Peroxide Capsule Self-Assembly in Slow Motion. Chem. Eur. J. 2019, 25 (24), 6087-6091. (35) Gaziev, S.; Gorshkov, N.; Mashirov, L.; Suglobov, D., Photostimulated oxygen exchange and photochemical properties of uranyl ions. Inorg. Chim. Acta 1987, 139 (1-2), 345-351. (36) Diwu, J.; Wang, S.; Good, J. J.; DiStefano, V. H.; Albrecht-Schmitt, T. E., Deviation between the chemistry of Ce (IV) and Pu (IV) and routes to ordered and disordered heterobimetallic 4f/5f and 5f/5f phosphonates. Inorg. Chem. 2011, 50 (11), 4842-4850. (37) APEX3 Bruker AXS Inc., Madison, WI, USA, 2016. (38) Sheldrick, G. i., Program for empirical absorption correction of area detector data. SADABS 1996. (39) Sheldrick, G. M., Crystal structure refinement with SHELXL. Acta Cryst. C 2015, 71 (1), 38. (40) Burns, P. C.; Ewing, R. C.; Hawthorne, F. C., The crystal chemistry of hexavalent uranium: polyhedron geometries, bond-valence parameters, and polymerization of polyhedra. Can. Mineral. 1997, 35, 1551-1570. (41) Brown, I.; Altermatt, D., Bond‐valence parameters obtained from a systematic analysis of the Inorganic Crystal Structure Database. Acta Cryst. B 1985, 41 (4), 244-247. (42) Pelta, M. D.; Morris, G. A.; Stchedroff, M. J.; Hammond, S. J., A one-shot sequence for high-resolution diffusion-ordered spectroscopy. Magn Reson Chem. 2002, 40 (13), S147-S152. (43) Sigmon, G. E.; Weaver, B.; Kubatko, K.-A.; Burns, P. C., Crown and Bowl-Shaped Clusters of Uranyl Polyhedra. Inorg. Chem. 2009, 48 (23), 10907-10909. (44) Sigmon, G. E.; Burns, P. C., Rapid self-assembly of uranyl polyhedra into crown clusters. J. Am. Chem. Soc. 2011, 133 (24), 9137-9139. (45) Roesky, H. W.; Andruh, M., The interplay of coordinative, hydrogen bonding and π–π stacking interactions in sustaining supramolecular solid-state architectures.: A study case of bis(4-pyridyl)- and bis(4-pyridyl-N-oxide) tectons. Coord Chem Rev. 2003, 236 (1), 91-119.

3

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

(46) Tsuzuki, S.; Honda, K.; Uchimaru, T.; Mikami, M.; Tanabe, K., Origin of Attraction and Directionality of the π/π Interaction:  Model Chemistry Calculations of Benzene Dimer Interaction. J. Am. Chem. Soc. 2002, 124 (1), 104-112. (47) Gresley, N. M.; Griffith, W. P.; Laemmel, A. C.; Nogueira, H. I. S.; Parkin, B. C., Studies on polyoxo and polyperoxo-metalates part 5: Peroxide-catalysed oxidations with heteropolyperoxo-tungstates and -molybdates. J. Mol. Catal. A: Chem. 1997, 117 (1), 185-198. (48) Bastians, S.; Crump, G.; Griffith, W. P.; Withnall, R., Raspite and studtite: Raman spectra of two unique minerals. J. Raman Spectrosc. 2004, 35 (8‐9), 726-731. (49) McGrail, B. T.; Sigmon, G. E.; Jouffret, L. J.; Andrews, C. R.; Burns, P. C., Raman Spectroscopic and ESI-MS Characterization of Uranyl Peroxide Cage Clusters. Inorg. Chem. 2014, 53 (3), 1562-1569. (50) Liao, Z.; Deb, T.; Nyman, M., Elucidating Self-Assembly Mechanisms of Uranyl–Peroxide Capsules from Monomers. Inorg. Chem. 2014, 53 (19), 10506-10513. (51) Dembowski, M.; Bernales, V.; Qiu, J.; Hickam, S.; Gaspar, G.; Gagliardi, L.; Burns, P. C., Computationally-Guided Assignment of Unexpected Signals in the Raman Spectra of Uranyl Triperoxide Complexes. Inorg. Chem. 2017, 56 (3), 1574-1580. (52) Bartlett, J. R.; Cooney, R. P., On the determination of uraniumoxygen bond lengths in dioxouranium(VI) compounds by Raman spectroscopy. J. Mol. Struct. 1989, 193, 295-300. (53) Nguyen Trung, C.; Begun, G. M.; Palmer, D. A., Aqueous uranium complexes. 2. Raman spectroscopic study of the complex formation of the dioxouranium(VI) ion with a variety of inorganic and organic ligands. Inorg. Chem. 1992, 31 (25), 5280-5287. (54) Lobeck, H. L.; Traustason, H.; Julien, P. A.; FitzPatrick, J. R.; Mana, S.; Szymanowski, J. E.; Burns, P. C., In situ Raman spectroscopy of uranyl peroxide nanoscale cage clusters under hydrothermal conditions. Dalton Trans. 2019, 48, 7755-7765. (55) Alam, T. M.; Liao, Z.; Zakharov, L. N.; Nyman, M., Solid-State Dynamics of Uranyl Polyoxometalates. Chem. Eur. J. 2014, 20 (27), 8302-8307. (56) Adelani, P. O.; Ozga, M.; Wallace, C. M.; Qiu, J.; Szymanowski, J. E. S.; Sigmon, G. E.; Burns, P. C., Hybrid Uranyl-Carboxyphosphonate Cage Clusters. Inorg. Chem. 2013, 52 (13), 7673-7679.

4

ACS Paragon Plus Environment

Page 26 of 27

Page 27 of 27 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

For Table of Contents Only

5

ACS Paragon Plus Environment