Positive and Negative Two-Dimensional Thermal Expansion via


Positive and Negative Two-Dimensional Thermal Expansion via...

0 downloads 87 Views 4MB Size

Article Cite This: Inorg. Chem. XXXX, XXX, XXX−XXX

pubs.acs.org/IC

Positive and Negative Two-Dimensional Thermal Expansion via Relaxation of Node Distortions Ryo Ohtani,*,† Riho Yamamoto,† Takuya Aoyama,‡ Arnaud Grosjean,§ Masaaki Nakamura,† Jack K. Clegg,§ and Shinya Hayami*,†,⊥

Inorg. Chem. Downloaded from pubs.acs.org by UNIV OF SOUTH DAKOTA on 09/06/18. For personal use only.



Department of Chemistry, Graduate School of Science and Technology and⊥Institute of Pulsed Power Science, Kumamoto University, 2-39-1 Kurokami, Chuo-ku, Kumamoto 860-8555 Japan ‡ Department of Physics, Graduate School of Science, Tohoku University, 6-3, Aramaki Aza-Aoba, Aoba-ku, Sendai, Miyagi 980-8578, Japan § School of Chemistry and Molecular Biosciences, The University of Queensland, St. Lucia, Queensland 4072 Australia S Supporting Information *

ABSTRACT: The ability to tune physical properties is attractive for the development of new materials for myriad applications. Understanding and controlling the structural dynamics in complicated network structures like coordination polymers (CPs) is particularly challenging. We report a series of two-dimensional CPs [Mn(salen)]2[M(CN)4]·xH2O (M = Pt (1), PtI2 (2), and MnN (3)) incorporating zigzag cyanonetwork layers that display composition-dependent anisotropic thermal expansion properties. Variable-temperature singlecrystal X-ray structural analyses demonstrated that the thermal expansion behavior is caused by double structural distortions involving [Mn(salen)]+ units incorporated into the zigzag layers. Thermal relaxations produce structural transformations resulting in positive thermal expansion for 2·H2O and negative thermal expansion for 3. In the case of 1·H2O, the relaxation does not occur and zero thermal expansion results in the plane between 200 to 380 K. The present study proposes a new strategy based on structural distortions in coordination networks to control thermal responsivities of frameworks.



INTRODUCTION

Three-dimensional (3D) CP frameworks as well as onedimensional (1D) and two-dimensional (2D) CPs have demonstrated anomalous mechanical properties such as colossal, anisotropic, or ZTE in individual materials.6,13,29−48 For example, 3D wine-rack frameworks show negative linear compressibility and NTE based on a topological mechanism involving hinge like motions.34−36 2D layers consisting of zigzag structures also shown anisotropic thermal expansion incorporating different behavior between in inter- and intralayers.31,37,38 Although 2D-CPs are expected to be suitable for the development of anisotropic structural properties via modifications to layer structures, tunable 2D thermal expansions has not been demonstrated thus far to the best of our knowledge. To understand how to tune such behavior in coordination polymers, it is required we determine the structural origins for thermal responsivities and molecular motion of the building blocks, as they contribute predominantly to the thermal expansion rather than vibrations. Such insights give further opportunities to find novel methodologies to control physical properties, together with the fundamental

Tunable mechanical properties such as thermal expansion, flexibility, and compressibility in solid-state materials are of fundamental interest and are attractive for technological applications.1−14 Unconventional negative thermal expansion (NTE) materials have been widely investigated, because a mainstream approach for the design of new materials, including those that display zero thermal expansion (ZTE), has often involved the combination of known positive thermal expansion (PTE) and NTE moieties. Many materials such as oxides,1,15 nitrides,16 ReO3-type compounds,2,3,17 metal− organic frameworks (or coordination polymers),18,19 metal cyanides,12,20−23 and metamaterials24 exhibit NTE which arises through low-energy transverse vibrations. Recently, tunable thermal expansion has been shown in MZrF6 (M = Ca, Mn, Fe, Co, Ni, and Zn) where flexible atomic linkages between the metal and fluorine ions result in different behavior for each compound.3 Tuning such behavior in more complex materials including organic−inorganic hybrid frameworks like coordination polymers (CPs)25−28 is less straightforward. In particular, the development of ZTE materials capable of switching between regions of PTE and NTE is a continuing challenge.1−5 © XXXX American Chemical Society

Received: June 10, 2018

A

DOI: 10.1021/acs.inorgchem.8b01617 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 1. Structures of two-dimensional cyano-bridged coordination polymers incorporating tetracyano-metalate units and their compositiondependent two-dimensional thermal expansion behavior.

Figure 2. Crystal structures of (a) 1·H2O, (b) 2·H2O, and (c) 3. Layer structures in ab plane (left), and stacking structures along c axis (center). Zigzag angles in layer structures are difined as the angles between respective tetracyanometallate units (right). Color code: pink (Mn), red (O), blue (N), and gray (C). H atoms are omitted for clarity. [Mn(salen)]2[Pt(CN)4]·H2O (1·H2O). A solution of [Mn(salen)]Cl (94.5 mg, 0.265 mmol) in H2O (100 mL) added to a solution of K2[Pt(CN)4]·3H2O (50.0 mg, 0.116 mmol) in MeOH (100 mL) at room temperature. Single crystals of 1·H2O were obtained after leaving to stand for 3 days. Powder samples of 1·H2O were collected as the right-brown precipitation by suction filtration after stirring for 24 h, and washed with water and methanol, and dried under vacuum. Yield 32.3%. Anal. Found (calcd) for C36H34Mn2N8O8Pt (995.61): C 43.39 (43.39); H 3.41 (3.41); N 11.25 (11.24). [Mn(salen)]2[Pt(CN)4(I)2]·H2O (2·H2O). A solution of [Mn(salen)]Cl (94.5 mg, 0.265 mmol) in H2O (100 mL) was added to a solution of K2[Pt(CN)4]·3H2O (50.0 mg, 0.116 mmol) and I2 (127 mg, 0.500 mmol) in MeOH (100 mL) at room temperature. Single crystals of 2· H2O were obtained after leaving to stand for 3 days. Powder samples of 2·H2O were collected as the right-brown precipitation by suction filtration after stirring for 24 h, washed with water and methanol, and dried under vacuum. Yield 40.5%. Anal. Found (calcd) for

views understanding the flexibility and the origin of thermal responsivities of complicated CP frameworks. Here we report a series of new flexible two-dimensional cyano-bridged CPs [Mn(salen)]2[M(CN)4]·xH2O (M = Pt (1), PtI2 (2), and MnN (3)) that show compositiondependent in plane ZTE, PTE, and NTE properties (Figure 1). We show that the TE behavior of framework originated from two intramolecular structural distortions of the [Mn(salen)]+ units.



EXPERIMENTAL SECTION

Synthesis. All reagents were commercially available and used without further purification. [Mn(salen)]Cl and (PPh4)2[Mn(N)(CN)4]·2H2O were synthesized according to the method described previously.49,50 B

DOI: 10.1021/acs.inorgchem.8b01617 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 3. (a) Thermal variation of a and b axes parameters. Temperature dependences of lattice constant changes relative to the 150 K values of (b) a and b axes, and (c) c axis. (d) Thermal variation in the angles between tetracyanometallate units in the zigzag layers. 1·H2O (black), 2·H2O (red), and 3 (blue). The standard deviation values are smaller than the plotted symbols.



C36H36I2Mn2N8O8Pt (1236.83): C 35.87 (35.59); H 2.84 (2.80); N 9.00 (9.22). [Mn(salen)]2[MnN(CN)4]·H2O (3·H2O). A solution of [Mn(salen)]Cl (20 mg, 0.055 mmol) in H2O (25 mL) added to a solution of K2[Mn(N)(CN)4] (7 mg, 0.028 mmol) in H2O (25 mL) at room temperature. Single crystals of 3·H2O were obtained after leaving to stand for 3 days. Physical Measurements. Single-crystal X-ray data for compounds were recorded on a Rigaku/MSC Saturn CCD diffractometer with confocal monochromated Mo−Kα (= 0.7107 Å) and processed by using Rigaku/CrystalClear software. The structures were solved by direct methods (Sir 2004) and refined by full-matrix least-squares refinement using the SHELXL-2013 computer program. The hydrogen atoms were refined geometrically by using a riding model. Thermogravimetric analysis was performed at 10 K min−1 using a Rigaku Instrument Thermo plus TG 8120 in a nitrogen atmosphere. Elemental analyses (C,H,N) were carried out on a J-SCIENCE LAB JM10 analyzer at the Instrumental Analysis Centre of Kumamoto University. Variable temperature powder X-ray diffraction patterns for 1 were obtained in SAGA-LS. The powder samples with homogeneous granularity sealed in a glass capillary of 0.5 mm internal diameter. Synchrotron radiation (SR) powder diffraction patterns were recorded on an imaging plate (IP) of the large Debye− Scherrer camera installed at BL1551 in SAGA-LS. The size of IP is 200 × 400 mm2, which covers up to 72 deg. in 2θ. The pixel size of IP is 50 × 50 μm2. The radius of the camera is 286.5 mm, which correspond to 0.01 deg. in 2θ. The energy of SR was 12.4 keV (wavelength λ = 1.000 Å). The sample temperatures were controlled by a dry dinitrogen flow using a Rigaku GN2 apparatus. The temperature dependence of the lattice constant of 1 was calculated from the results of powder X-ray diffraction patterns using a RIETANFP software.52 Above a structural transition temperature, the results for single-crystal 1·H2O were referred. Below the structural transition temperature, the orthorhombic space group is chosen from the maximal-subgroup of P4/ncc to determine the Miller Indices.

RESULTS AND DISCUSSION Preparation and Crystal Structures. Single crystals of [Mn(salen)]2[Pt(CN)4]·H2O (1·H2O) were prepared by mixing methanol solutions of [Mn(salen)]Cl and K2[Pt(CN)4]·3H2O in a 2:1 ratio.53,54 At 150 K 1·H2O crystallizes in a tetragonal (P4/ncc). Undulating zigzag layers are constructed from two-connecting [Mn(salen)]+ units linked through the nitrogen atoms of the cyanide groups of fourconnecting [Pt(CN)4]2− units (Figure 2(a)). The infinite 2D layers propagate in the ab plane and are stacked along the c axis. The angles between [Pt(CN)4]2− units in the zigzag layers are 166.78°. H2O solvent molecules are accommodated in the center of the [Pt(CN)4]2− grids. Thermogravimetric analysis (TGA) of powder samples of 1·H2O showed that the lattice H2O solvent molecules are lost above 450 K (Figure S1). Single crystals of [Mn(salen)]2[PtI2(CN)4]·H2O (2·H2O) were prepared by mixing methanol solutions of [Mn(salen)]Cl, K2[Pt(CN)4]·3H2O and I2 in a 2:1:5 ratio. 2·H2O consists of a similar layer structure to that of 1·H2O, with smaller angles (157.45°) between [PtI2(CN)4]2− units (Figure 2b). The lattice H2O solvent molecules of 2·H2O are lost above 420 K (Figure S1). Single crystals of [Mn(salen)]2[MnN(CN)4]·H2O (3·H2O) were prepared by mixing aqueous solutions of [Mn(salen)]Cl and K2[MnN(CN)4]49,50 in a 2:1 ratio. As heating these crystals, the lattice water molecules are removed gradually under 350 K (Figure 3a), producing the desolvated form, 3. The angles between [MnN(CN)4]2− units in the zigzag layers of 3 are 158.18° (Figure 2c). Thermal Expansion Behavior. The thermal expansion behavior of each compound was investigated by variabletemperature single-crystal X-ray structural analyses (VTSCXRD) (Figure 3, Figure S2, and Tables S1−S4). The C

DOI: 10.1021/acs.inorgchem.8b01617 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

their expansion (Figure 3d). This structural transformation of 2D cyano-network layers occur with the rotations of [MnO2N4] cores in [Mn(salen)]+ units. Here, one [Mn(salen)]+ unit (selected as shown in Figure 4−6) is focused for investigating the structural transformation, whereas its neighboring [Mn(salen)]+ units do exhibit transformations by the opposite trend due to the symmetric layer structures (Figure 7). Cyano-networks of 2·H2O are distorted in counterclockwise directions, as shown in Figure 4b. As increasing temperatures, the [MnO2N4] cores rotated decreasing the distortions in cyano-networks resulted in PTE (Figure 4c, d). Moreover, at the same time, transformations of molecular octahedral geometry of [Mn(salen)] units were observed (Figure 5). The molecular structure of [Mn(salen)]+ units are distorted clockwise, in which the angles of CN-Mnbenzene ring are 94.39° and 85.12° at 150 K (Figure 5c). These angles approarch 90°, that is closer to the ideal octahedral geometry, by the rotation of salen ligands as increasing temperatures. These results demonstrate that [MnO2N4] cores and salen ligands rotate in the opposite directions to relax their structural mismatches in [Mn(salen)]+ units; [MnO2N4] cores rotate clockwise while salen ligands rotate counterclockwise in 2·H2O (Figure 6a and 7b and Figure S3). In the cases of 3 with NTE, the angles of the zigzag layers decrease with increasing temperatures to produce a contraction in the layers, demonstrating the opposite behavior to 2·H2O (Figure 3d). Smaller distortions in cyano-networks are observed in 3 than 2·H2O (Figure 4b). The molecular structure of [Mn(salen)]+ units are distorted counterclockwise,

thermal expansion coefficients are given in Table 1. 1·H2O shows little change of the a and b axes above 200 K (αa,b = Table 1. Calculated Thermal Expansion Coefficients (α) along a, b Axes and c Axis for 1·H2O, 2·H2O, and 3 1·H2O (200 K−380 K) 2·H2O (150 K−370 K) 3 (150 K−400 K)

αa,b (MK−1)

αc (MK−1)

−1.49 +46.4 −15.9

+55.7 +38.9 +103

−1.49 MK−1; M = 1 × 10−6) producing zero 2D thermal expansion within the layers, whereas the interlayer spacing continues to increase (αc = +55.7 MK−1). 2·H2O, on the other hand, shows PTE from 150 to 370 K (αa,b = +46.4 MK−1 and αc = +38.9 MK−1). 3 shows 2D negative TE (NTE) within the layers (αa,b = −15.9 MK−1), whereas the interlayer spacing expands (αc = +103 MK−1). Each of 1·H2O, 2·H2O, and 3 therefore show ZTE, PTE, and NTE within the 2D layers, respectively. The volumetric thermal expansion coefficients of 1·H2O, 2·H2O, and 3 are 52.8, 133, and 70.5 MK−1, respectively (Figure S2). Mechanism of Different Thermal Expansions of 2D Layers. The diverse thermal expansion behavior of the three compounds results from structural changes of zigzag layers. The thermal variation of the angles between [M(CN)4]2− (M = Pt for 1·H2O, PtI2 for 2·H2O, and MnN for 3) units in the zigzag layers are shown in Figure 3d. In 2·H2O with PTE, the angles increase with increasing temperature up to 370 K, producing flatter layer structures and

Figure 4. (a) Schematic views displaying structural distortions in cyano-network layers. (b) Schematic views displaying cyano-network layers and their distortion angles of 1·H2O, 2·H2O, and 3. The distortion angles are indicated by blue lines. (c) Thermal variation of the distortion angles in cyano-network layers. (d) Schematic views displaying structural transformations of cyano-network layers for 2·H2O and 3 by heating. D

DOI: 10.1021/acs.inorgchem.8b01617 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 5. (a) Schematic views displaying structural distortions, counterclockwise (red) and clockwise (blue), of a [Mn(salen)]+ unit. Thermal variation of angles in CN−Mn--benzene rings in [Mn(salen)]+ units and schematic views displaying structural transformations for (b) 1·H2O, (c) 2·H2O, and (d) 3. Black dashed arrows in c and d indicate directions of the structural transformations by heating. 2·H2O shows a structrual transformation to decrease the red angles and increase the black angles, whereas the transformation of 3 increases the red angles and decrease the black angles.

and Figure S3). No thermal response of distortions in [Mn(salen)]+ units in 1 is likely because both [MnO2N4] cores and salen ligands are distorted counterclockwise as the same “directions” (Figures 4 and 5). This situation in 1·H2O prevents two structural distortions from thermal responses of rotations to decrease the mismatches in [Mn(salen)]+ units (Figure 7a). The zigzag layers in 1·H2O exhibit a slight contraction (αa,b = −1.49 MK−1) above 200 K without decreasing the local distortions in [Mn(salen)]+ units (Table 1). This behavior might arise from the transverse stretching mode of cyanides of [Pt(CN)4]2−, which has been observed in a similar 2D-CP of FePt(CN)4.55,56 The effect of lattice waters in 1·H2O on its TE behavior has been investigated by VT-powder XRD between 100 and 440 K in the Kyushu synchrotron radiation source (SAGA-LS) for powder 1 that was obtained by a thermal treatment for 1·H2O at 550 K (Figure 8). 1 exhibits a

in which the angles of CN−Mn−benzene ring are 75.28 and 104.31° for 3 at 150 K (Figure 5d). As increasing temperatures, [MnO2N4] cores rotate counterclockwise (Figure 4d), whereas salen ligands rotate clockwise (Figure 5d) to decrease the mismatch angles in [Mn(salen)]+ units, producing the contraction of the layers (Figure 6b, c and Figure S3). Along this NTE behavior with the rotation of [MnO2N4] cores, distortions in cyano-networks are increased slightly. This result indicates that the decrease of the mismach angles in [Mn(salen)]+ units predominates thermal responses (Figure 7c). On the other hand, 1·H2O with ZTE shows no thermal variation of the angles in the zigzag layers (Figure 3d). Although 1·H2O incorporates similar distortions in [Mn(salen)]+ units (Figures 4 and 5) where the mismatch angle is 6.7°, they are not decreased without rotations of [MnO2N4] cores and salen ligands as increasing temperatures (Figure 6c E

DOI: 10.1021/acs.inorgchem.8b01617 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 7. Schematic views displaying structural transformations for (a) ZTE, (b) PTE, and (c) NTE in the plane. Differences in the directions of each distortion are expressed by different colors. In ZTE, both [MnO2N4] cores and salen ligands are distorted in the same direction. In PTE, on the other hand, [MnO2N4] cores and salen ligands are distorted in the opposite direction.

Figure 6. Schematic views displaying structural transformations decreasing the mismatch angles in [Mn(salen)]+ units for (a) 2· H2O and (b) 3. The mismatch angles are indicated by blue lines. Blue arrows indicate directions of the structural transformations by heating. (c) Thermal variation in the mismatch angles as increasing temperatures for 1·H2O, 2·H2O, and 3.

metalate units [Pt(CN)4]2−, [PtI2(CN)4]2−, and [MnN(CN)4]2−. Planar-shaped [Pt(CN)4]2− and [PtI2(CN)4]2− units incorporating cyanides angles of 180° unambiguously produce [MnO2N4] cores distorted counterclockwise (Figure 4b). Umbrella-shaped [MnN(CN)4]2− incorporating cyanides angles of 160° corresponding to angles of zigzag layers give rise to smaller distortions in cyano-networks (Figure 4b). On the other hand, distortions of molecular octahedral geometries of [Mn(salen)]+ units are caused by the structures of axial sites of tetracyano-metalate units. In the case of [Pt(CN)4]2− and [MnN(CN)4]2− incorporating open metal axial sites, salen ligands are distorted in the direction to fill the open spaces (counterclockwise) (Figure 5b, d). [PtI2(CN)4]2− units, on the other hand, produce a steric hindrance to neighboring benzen

structural transformation at 200 K between orthorhombic and tetragonal (P4/ncc system) phases. Above 200 K, 1 shows NTE in the plane (αa,b = −15.5 MK−1, αc = 82.0 MK−1; 200− 440 K). The volumetric thermal expansion coefficient of 1 is 50.5 MK−1. This result indicates that lattice waters might damp the contraction of zigzag layers caused by the transverse stretching mode of cyanides in 1·H2O, resulted in the observed ZTE. Structural distortions involving [MnO2N4] cores in cyanonetworks are caused by angles between cyanides in tetracyanoF

DOI: 10.1021/acs.inorgchem.8b01617 Inorg. Chem. XXXX, XXX, XXX−XXX

Inorganic Chemistry



Article

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.8b01617. Table of crystal parameters, TGA results, thermal variation of volumes (PDF) Accession Codes

CCDC 1840107−1840124 and 1848131−1848134 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/ data_request/cif, or by emailing [email protected]. uk, or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.



AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected] (R.O.). *E-mail: [email protected] (S.H.). ORCID

Ryo Ohtani: 0000-0003-4840-3338 Jack K. Clegg: 0000-0002-7140-5596 Shinya Hayami: 0000-0001-8392-2382 Author Contributions

The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was supported by JSPS Grant-in-Aid for Young Scientists (B) JP18K14245 and JSPS Grant-in-Aid for Scientific Research on Innovative Areas (Dynamical Ordering & Integrated Functions JP16H00777, Mixed Anion JP17H05485, JP17H05474). This work was also supported by KAKENHI Grant-in-Aid for Scientific Research (A) JP17H01200. This work was partially supported by the Cooperative Research Program of ‘Network Joint Research Centre for Materials and Devices’, and Kyushu Synchrotron Light Research Center 1606042F. We acknowledge the support of the Australian Research Council.

Figure 8. (a) VT-PXRD patterns for 1. (b) Thermal variation in cell parameters of 1 (filled circles) and 1·H2O (open circles).

rings in salen ligands, giving rise to the distortion in the opposite direction (clockwise) to the other cases (Figure 5c).





CONCLUSION We demonstrated anisotropic zero, positive, and negative thermal expansions of 2D-CPs consisting of zigzag layers. Intramolecular double distortions and their thermal relaxation phenomena in the building blocks are responsible for the TE behavior. Structural distortions in frameworks produce thermal responsivity involving structural transformations, and they would have an active role in the flexibilities of stimuli responsive CPs. By means of the design preventing thermal relaxation of distortions, zero anisotropic thermal expansion of 2D-CPs resulted. These results will lead not only to the better understanding of their structural dynamics and mechanical properties of CP materials incorporating complicated structures but also to the development of future devices such as molecular actuators utilizing ZTE in-plane and out-of-plane flexibility of 2D materials.

REFERENCES

(1) Chen, J.; Hu, L.; Deng, J. X.; Xing, X. R. Negative thermal expansion in functional materials: controllable thermal expansion by chemical modifications. Chem. Soc. Rev. 2015, 44, 3522−3567. (2) Chen, J.; Gao, Q. L.; Sanson, A.; Jiang, X. X.; Huang, Q. Z.; Carnera, A.; Rodriguez, C. G.; Olivi, L.; Wang, L.; Hu, L.; Lin, K.; Ren, Y.; Lin, Z. S.; Wang, C.; Gu, L.; Deng, J. X.; Attfield, J. P.; Xing, X. R. Tunable thermal expansion in framework materials through redox intercalation. Nat. Commun. 2017, 8, 14441. (3) Hu, L.; Chen, J.; Xu, J. L.; Wang, N.; Han, F.; Ren, Y.; Pan, Z.; Rong, Y. C.; Huang, R. J.; Deng, J. X.; Li, L. F.; Xing, X. R. Atomic Linkage Flexibility Tuned Isotropic Negative, Zero, and Positive Thermal Expansion in MZrF6 (M = Ca, Mn, Fe, Co, Ni, and Zn). J. Am. Chem. Soc. 2016, 138, 14530−14533. (4) Tallentire, S. E.; Child, F.; Fall, I.; Vella-Zarb, L.; Evans, I. R.; Tucker, M. G.; Keen, D. A.; Wilson, C.; Evans, J. S. O. Systematic and Controllable Negative, Zero, and Positive Thermal Expansion in Cubic Zr1‑xSnxMo2O8. J. Am. Chem. Soc. 2013, 135, 12849−12856. G

DOI: 10.1021/acs.inorgchem.8b01617 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

analogues MIIPtIV(CN)6·x{H2O} (0 ≤ x ≤ 2; M = Zn, Cd). J. Am. Chem. Soc. 2005, 127, 17980−17981. (24) Wu, L. L.; Li, B.; Zhou, J. Isotropic Negative Thermal Expansion Metamaterials. ACS Appl. Mater. Interfaces 2016, 8, 17721−17727. (25) Horike, S.; Shimomura, S.; Kitagawa, S. Soft porous crystals. Nat. Chem. 2009, 1, 695−704. (26) Ferey, G. Hybrid porous solids: past, present, future. Chem. Soc. Rev. 2008, 37, 191−214. (27) Kitagawa, S.; Kitaura, R.; Noro, S. Functional porous coordination polymers. Angew. Chem., Int. Ed. 2004, 43, 2334−2375. (28) Yaghi, O. M.; O’Keeffe, M.; Ockwig, N. W.; Chae, H. K.; Eddaoudi, M.; Kim, J. Reticular synthesis and the design of new materials. Nature 2003, 423, 705−714. (29) Mullaney, B. R.; Goux-Capes, L.; Price, D. J.; Chastanet, G.; Letard, J. F.; Kepert, C. J. Spin crossover-induced colossal positive and negative thermal expansion in a nanoporous coordination framework material. Nat. Commun. 2017, 8, 1053. (30) Hu, J. X.; Xu, Y.; Meng, Y. S.; Zhao, L.; Hayami, S.; Sato, O.; Liu, T. A Material Showing Colossal Positive and Negative Volumetric Thermal Expansion with Hysteretic Magnetic Transition. Angew. Chem., Int. Ed. 2017, 56, 13052−13055. (31) Ohtani, R.; Grosjean, A.; Ishikawa, R.; Yamamoto, R.; Nakamura, M.; Clegg, J. K.; Hayami, S. Zero in-Plane Thermal Expansion in Guest-Tunable 2D Coordination Polymers. Inorg. Chem. 2017, 56, 6225−6233. (32) Lama, P.; Alimi, L. O.; Das, R. K.; Barbour, L. J. Hydrationdependent anomalous thermal expansion behaviour in a coordination polymer. Chem. Commun. 2016, 52, 3231−3234. (33) Das, R. K.; Aggarwal, H.; Barbour, L. J. Anomalous Anisotropic Thermal Expansion in a One-Dimensional Coordination Polymer Driven by Conformational Flexibility. Inorg. Chem. 2015, 54, 8171− 8173. (34) Zeng, Q.; Wang, K.; Zou, B. Large Negative Linear Compressibility in InH(BDC)(2) from Framework Hinging. J. Am. Chem. Soc. 2017, 139, 15648−15651. (35) Cai, W. Z.; Katrusiak, A. Giant negative linear compression positively coupled to massive thermal expansion in a metal-organic framework. Nat. Commun. 2014, 5, 4337. (36) Zhang, L.; Kuang, X.; Wu, X.; Yang, W.; Lu, C. Supramolecular interactions induced hinge-like motion of a metal-organic framework accompanied by anisotropic thermal expansion. Dalton Trans 2014, 43, 7146−7152. (37) Hodgson, S. A.; Adamson, J.; Hunt, S. J.; Cliffe, M. J.; Cairns, A. B.; Thompson, A. L.; Tucker, M. G.; Funnell, N. P.; Goodwin, A. L. Negative area compressibility in silver(I) tricyanomethanide. Chem. Commun. 2014, 50, 5264−5266. (38) Sergeenko, A. S.; Ovens, J. S.; Leznoff, D. B. Designing anisotropic cyanometallate coordination polymers with unidirectional thermal expansion (TE): 2D zero and 1D colossal positive TE. Chem. Commun. 2018, 54, 1599−1602. (39) Zhou, H.-L.; Lin, R.-B.; He, C.-T.; Zhang, Y.-B.; Feng, N.; Wang, Q.; Deng, F.; Zhang, J.-P.; Chen, X.-M. Direct visualization of a guest-triggered crystal deformation based on a flexible ultramicroporous framework. Nat. Commun. 2013, 4, 2534. (40) Wei, Y. S.; Chen, K. J.; Liao, P. Q.; Zhu, B. Y.; Lin, R. B.; Zhou, H. L.; Wang, B. Y.; Xue, W.; Zhang, J. P.; Chen, X. M. Turning on the flexibility of isoreticular porous coordination frameworks for drastically tunable framework breathing and thermal expansion. Chem. Sci. 2013, 4, 1539−1546. (41) Henke, S.; Schneemann, A.; Fischer, R. A. Massive Anisotropic Thermal Expansion and Thermo-Responsive Breathing in MetalOrganic Frameworks Modulated by Linker Functionalization. Adv. Funct. Mater. 2013, 23, 5990−5996. (42) Keene, T. D.; Murphy, M. J.; Price, J. R.; Price, D. J.; Kepert, C. J. A new modification of an old framework: Hofmann layers with unusual tetracyanidometallate groups. Dalton Trans 2011, 40, 11621− 11628.

(5) Gao, Q. L.; Chen, J.; Sun, Q.; Chang, D. H.; Huang, Q. Z.; Wu, H.; Sanson, A.; Milazzo, R.; Zhu, H.; Li, Q.; Liu, Z. N.; Deng, J. X.; Xing, X. R. Switching Between Giant Positive and Negative Thermal Expansions of a YFe(CN)6 based Prussian Blue Analogue Induced by Guest Species. Angew. Chem., Int. Ed. 2017, 56, 9023−9028. (6) Tan, J. C.; Cheetham, A. K. Mechanical properties of hybrid inorganic-organic framework materials: establishing fundamental structure-property relationships. Chem. Soc. Rev. 2011, 40, 1059− 1080. (7) Boatti, E.; Vasios, N.; Bertoldi, K. Origami Metamaterials for Tunable Thermal Expansion. Adv. Mater. 2017, 29, 1700360. (8) Wang, T.; Xu, J. L.; Hu, L.; Wang, W.; Huang, R. J.; Han, F.; Pan, Z.; Deng, J. X.; Ren, Y.; Li, L. F.; Chen, J.; Xing, X. R. Tunable thermal expansion and magnetism in Zr-doped ScF3. Appl. Phys. Lett. 2016, 109, 181901. (9) Xu, H.; Pasini, D. Structurally Efficient Three-dimensional Metamaterials with Controllable Thermal Expansion. Sci. Rep. 2016, 6, 34924. (10) Duyker, S. G.; Peterson, V. K.; Kearley, G. J.; Studer, A. J.; Kepert, C. J. Extreme compressibility in LnFe(CN)6 coordination framework materials via molecular gears and torsion springs. Nat. Chem. 2016, 8, 270−275. (11) Chapman, K. W.; Chupas, P. J. Pressure enhancement of negative thermal expansion behavior and induced framework softening in zinc cyanide. J. Am. Chem. Soc. 2007, 129, 10090−10091. (12) Chapman, K. W.; Chupas, P. J.; Kepert, C. J. Compositional dependence of negative thermal expansion in the Prussian blue analogues MIIPtIV(CN)6 (M = Mn, Fe, Co, Ni, Cu, Zn, Cd). J. Am. Chem. Soc. 2006, 128, 7009−7014. (13) Grobler, I.; Smith, V. J.; Bhatt, P. M.; Herbert, S. A.; Barbour, L. J. Tunable Anisotropic Thermal Expansion of a Porous Zinc(II) Metal-Organic Framework. J. Am. Chem. Soc. 2013, 135, 6411−6414. (14) Worthy, A.; Grosjean, A.; Pfrunder, M. C.; Xu, Y.; Yan, C.; Edwards, G.; Clegg, J. K.; McMurtrie, J. C. Atomic resolution of structural changes in elastic crystals of copper(II) acetylacetonate. Nat. Chem. 2018, 10, 65−69. (15) Mary, T. A.; Evans, J. S. O.; Vogt, T.; Sleight, A. W. Negative thermal expansion from 0.3 to 1050 K in ZrW2O8. Science 1996, 272, 90−92. (16) Lin, J. C.; Tong, P.; Tong, W.; Lin, S.; Wang, B. S.; Song, W. H.; Zou, Y. M.; Sun, Y. P. Tunable negative thermal expansion related with the gradual evolution of antiferromagnetic ordering in antiperovskite manganese nitrides Ag1‑xNMn3+x (0 ≤ x ≤ 0.6). Appl. Phys. Lett. 2015, 106, 082405. (17) Hu, L.; Chen, J.; Fan, L. L.; Ren, Y.; Rong, Y. C.; Pan, Z.; Deng, J. X.; Yu, R. B.; Xing, X. R. Zero Thermal Expansion and Ferromagnetism in Cubic Sc1‑xMxF3 (M = Ga, Fe) over a Wide Temperature Range. J. Am. Chem. Soc. 2014, 136, 13566−13569. (18) Lock, N.; Christensen, M.; Kepert, C. J.; Iversen, B. B. Effect of gas pressure on negative thermal expansion in MOF-5. Chem. Commun. 2013, 49, 789−791. (19) Wu, Y.; Kobayashi, A.; Halder, G. J.; Peterson, V. K.; Chapman, K. W.; Lock, N.; Southon, P. D.; Kepert, C. J. Negative Thermal Expansion in the Metal-Organic Framework Material Cu3(1,3,5benzenetricarboxylate)2. Angew. Chem., Int. Ed. 2008, 47, 8929−8932. (20) Duyker, S. G.; Peterson, V. K.; Kearley, G. J.; Ramirez-Cuesta, A. J.; Kepert, C. J. Negative Thermal Expansion in LnCo(CN)6 (Ln = La, Pr, Sm, Ho, Lu, Y): Mechanisms and Compositional Trends. Angew. Chem., Int. Ed. 2013, 52, 5266−5270. (21) Phillips, A. E.; Goodwin, A. L.; Halder, G. J.; Southon, P. D.; Kepert, C. J. Nanoporosity and exceptional negative thermal expansion in single-network cadmium cyanide. Angew. Chem., Int. Ed. 2008, 47, 1396−1399. (22) Margadonna, S.; Prassides, K.; Fitch, A. N. Zero thermal expansion in a Prussian blue analogue. J. Am. Chem. Soc. 2004, 126, 15390−15391. (23) Goodwin, A. L.; Chapman, K. W.; Kepert, C. J. Guestdependent negative thermal expansion in nanoporous Prussian Blue H

DOI: 10.1021/acs.inorgchem.8b01617 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry (43) Goodwin, A. L.; Kennedy, B. J.; Kepert, C. J. Thermal Expansion Matching via Framework Flexibility in Zinc Dicyanometallates. J. Am. Chem. Soc. 2009, 131, 6334−6335. (44) Goodwin, A. L.; Keen, D. A.; Tucker, M. G.; Dove, M. T.; Peters, L.; Evans, J. S. O. Argentophilicity-dependent colossal thermal expansion in extended Prussian blue analogues. J. Am. Chem. Soc. 2008, 130, 9660−9661. (45) Goodwin, A. L.; Calleja, M.; Conterio, M. J.; Dove, M. T.; Evans, J. S. O.; Keen, D. A.; Peters, L.; Tucker, M. G. Colossal positive and negative thermal expansion in the framework material Ag3[Co(CN)6]. Science 2008, 319, 794−797. (46) Li, J.; Bi, W. H.; Ki, W.; Huang, X. Y.; Reddy, S. Nanostructured crystals: Unique hybrid semiconductors exhibiting nearly zero and tunable uniaxial thermal expansion behavior. J. Am. Chem. Soc. 2007, 129, 14140−14141. (47) Liu, Z. N.; Li, Q.; Zhu, H.; Lin, K.; Deng, J. X.; Chen, J.; Xing, X. R. 3D negative thermal expansion in orthorhombic MIL-68(In). Chem. Commun. 2018, 54, 5712−5715. (48) Liu, Z. N.; Gao, Q. L.; Chen, J.; Deng, J. X.; Lin, K.; Xing, X. R. Negative thermal expansion in molecular materials. Chem. Commun. 2018, 54, 5164−5176. (49) Bendix, J.; Meyer, K.; Weyhermuller, T.; Bill, E.; Metzler-Nolte, N.; Wieghardt, K. Nitridocyanometalates of CrV, MnV, and MnVI. Inorg. Chem. 1998, 37, 1767−1775. (50) Ohtani, R.; Kitamura, Y.; Hijikata, Y.; Nakamura, M.; Lindoy, L. F.; Hayami, S. Modulation of redox potentials utilizing the flexible coordination sphere of a penta-coordinate complex in the solid state. Dalton Trans 2017, 46, 3749−3754. (51) Okajima, T.; Chikaura, Y.; Tabata, M.; Hashimoto, H.; Soejima, Y.; Hara, K.; Hiramatsu, N. Design of beamline BL15 at the SAGA light source. Nucl. Instrum. Methods Phys. Res., Sect. B 2005, 238, 185−188. (52) Izumi, F.; Momma, K. Three-dimensional visualization in powder diffraction. Solid State Phenom. 2007, 130, 15−20. (53) Nemec, I.; Herchel, R.; Travnicek, Z. Pentacoordinate and Hexacoordinate Mn(III) Complexes of Tetradentate Schiff-Base Ligands Containing Tetracyanidoplatinate(II) Bridges and Revealing Uniaxial Magnetic Anisotropy. Molecules 2016, 21, 1681. (54) Yuan, A. H.; Shen, X. P.; Yu, K. B.; Lu, L. D.; Zhou, W.; Zhong, J. J. Synthesis, crystal structure and magnetic properties of a cyanobridged two-dimensional coordination polymer {[Fe(salen)]2[Ni(CN)4]}n. Acta Chim. Sinica 2003, 61, 1603−1607. (55) Goodwin, A. L.; Kepert, C. J. Negative thermal expansion and low-frequency modes in cyanide-bridged framework materials. Phys. Rev. B: Condens. Matter Mater. Phys. 2005, 71, 140301. (56) van der Veen, R. M.; Tissot, A.; Hauser, A.; Zewail, A. H. Unusual molecular material formed through irreversible transformation and revealed by 4D electron microscopy. Phys. Chem. Chem. Phys. 2013, 15, 7831−7838.

I

DOI: 10.1021/acs.inorgchem.8b01617 Inorg. Chem. XXXX, XXX, XXX−XXX