Axial Positive, Negative, and Zero Thermal Expansion in a Mixed

7 days ago - Zero thermal expansion (ZTE) occurs, when a material neither expands nor ... (57) Very recently, we have accounted for area NTE in a mixe...
0 downloads 0 Views 3MB Size
Communication Cite This: Cryst. Growth Des. XXXX, XXX, XXX−XXX

pubs.acs.org/crystal

Axial Positive, Negative, and Zero Thermal Expansion in a MixedMetal Mixed-Linker Coordination Compound: Role of 2D Layer in the Thermal Expansion Property Ashutosh Shrivastava and Dinabandhu Das* School of Physical Sciences, Jawaharlal Nehru University, New Delhi-110067, India

Downloaded via UNIV STRASBOURG on August 21, 2019 at 14:54:53 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

S Supporting Information *

ABSTRACT: Thermal expansion of a new mixed-metal mixed-linker metal organic framework (MOF) has been studied by variable temperature single crystal X-ray diffraction. Origin of axial positive, negative, and zero thermal expansion in the MOF has been elucidated by analyzing the crystal structures. Flexibility in coordination of ligands around the metal ions induces axial NTE and PTE in the 2D layers of the MOF. On the other hand, thickness of the 2D layers in the crystal structure remains unchanged along one of the principal axes resulting in rare ZTE. O···O Distances of O−H···O hydrogen bonding between the 2D layers in the crystal structure remain almost unaffected with increasing temperature.

M

For last three decades, tremendous growth in the synthesis and characterization of metal organic framework (MOF) materials has added new dimensions in materials science.41−51 Due to different chemical versatility as well as structural diversity, MOFs possess novel architecture that can be utilized for diverse applications in electronics,50 catalysis,48 drug delivery,51 and gas storage and separation.43,45,49 The different architectures of MOFs can be tuned by varying the flexibility of the organic ligands coordinated to the metal atoms that could produce a diverse framework structures possessing the desired properties for different applications.52,53 Owing to the topological diversity, MOFs can be used as a unique platform to study various unusual behaviors including thermomechanical properties. Anomalous thermal expansion has been observed recently in many MOFs. Omary et al. showed large PTE and NTE in a breathable fluorous MOF in the presence and absence of gas molecules.54 Recently, Fischer et al. showed a massive framework expansion possessing unprecedentedly large PTE and NTE in a functionalized MOF comprising mixed organic ligands.12 A mixed-metal mixed-linker MOF can be very useful to tune thermal expansion by varying either metal or ligand, or both metal and ligand coordinated with metals. Investigation of the thermal expansion property in mixed-metal mixed-ligand MOFs has been very rare.55,56 Colossal volumetric NTE and PTE has been observed in a mixed-metal MOF reported by Liu et al.35 Recently, Filhol et al. and others have described the colossal PTE and giant NTE of a mixed-linker MOF.57 Very recently, we have accounted for area NTE in a mixed-metal mixed-organic MOF.16 Although PTE, NTE, and ZTE have been reported by Barbour et al.,

echanical properties of materials induced by temperature have attracted significant attention of researchers in recent years.1 Generally, a material expands along all three axes upon heating, owing to increasing anharmonic vibration. This phenomenon is known as positive thermal expansion (PTE). However, some materials contract upon heating, termed negative thermal expansion (NTE). Zero thermal expansion (ZTE) occurs, when a material neither expands nor contracts with the change of temperature. Since the discovery of isotropic NTE of ZrW2O8 in 1996,2 significant developments have happened in the area of NTE materials, although it is restricted in certain classes of materials, viz., cyano-bridged metal complexes,3−5 Prussian blue analogues,6−8 and MOFs including some metal oxides.9−16 In addition, some organic compounds have shown an axial NTE property.17−24 A rare phenomenon, ZTE has been found in Invar, but few other materials.25−28 Materials with NTE and ZTE properties have potential applications in electronics, optical fiber systems, thermomechanical actuators and sensors, cookware, high precision optical devices, and aerospace engineering.29,30 ZTE or near-ZTE materials can be prepared as composites of PTE and NTE materials.27,28 Normally, the magnitude of the linear thermal expansion coefficient (α) for many materials is typically on the order of ∼20 MK−1. However, there are few materials reported with unusually high α. The number of the mechanism has been proposed to explain the unusual behaviors of the materials such as transverse vibration,2,3,5,31−33 phase transition,34−36 electronic effect,37 geometric variation,38,39 magnetostriction,40 and supramolecular interaction.14,16 NTE in organic molecules has been explained by tilting, sliding mechanism as well as steric hindrance.17−24 Takenaka has categorized all the mechanisms for NTE into three groups: flexible network, atomic radius contraction, and magnetovolume effect.30 © XXXX American Chemical Society

Received: July 5, 2019 Revised: August 6, 2019 Published: August 14, 2019 A

DOI: 10.1021/acs.cgd.9b00880 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Communication

Figure 1. (a) Asymmetric unit of 1; (b) SBU in the crystal structure of 1; (c) 1D linear array of SBUs propagating along the a axis; (d) 2D layers in the crystal structure of 1 viewed down the b axis. Hydrogen bonds are shown in the red dotted line. All the hydrogen atoms except the carboxylic hydrogen atoms are omitted for clarity. The 1D linear architecture in the crystal structure has been shown by the yellow shade.

within the temperature range of 160 K in a solvated MOF,58 the same has not been reported before in any mixed-metal mixed-linker framework structure in a wider range of temperature. Here, we report the thermal expansion property of a new mixed-metal mixed-organic 2D framework material comprising Co(II) and Zn(II) ions coordinated with two different organic linkers (i.e., benzene tricarboxylic acid and 4,4′-bipyridine). Variable temperature single crystal X-ray diffraction (VT-SCXRD) shows a combination of axial NTE, ZTE, and PTE within the temperature range from 100 to 350 K. The role of 2D layers in the crystal structure for the illucidation of axial NTE, ZTE, and PTE in the mixed-metal mixed-linker MOF is an imporatnt aspect of this study. Hydrothermal reaction of ZnCl2, CoCl2, benzene tricarboxylic acid (H3BTC), 4,4′-bipyridine (bpy), and KOH in distilled water at 180 °C produced purple color block-shaped single crystals (see Supporting Information for synthetic details). Structure determined by single crystal X-ray diffraction (SCXRD) experiment shows the formula unit of the

compound as [Co1Zn1(μ3-OH)1(HBTC)1(H2BTC)1(bpy)1.5] (1) observed in the asymmetric unit. The presence of Co and Zn ions was confirmed by SEM-EDAX experiment (see Figure S1 in Supporting Information). Compound 1 was crystallized in space group Pi̅. The moiety of the asymmetric unit of 1, consisting of one Co(II) ion, one Zn(II) ion, one of each of HBTC2−, H2BTC1−, and OH− ions, and one and half molecules of bpy (Figure 1a). This implies the presence of two symmetrically different bpy molecules in the crystal structure. One of the nitrogen atoms of the asymmetric bpy molecule coordinates with Zn(II) ions, and the other nitrogen atom of the same bpy remains uncoordinated with metal ions but participates in the formation of O−H···N hydrogen bonding, whereas both nitrogen atoms of the symmetric bpy molecules bind with Co(II) ions. Zn(II) ions are coordinated in tetrahedral geometry by binding with a nitrogen atom of the bpy and three oxygen atoms, out of which one is from HBTC2−, another from H2BTC−, and the third from the OH− ion. On the other hand, Co(II) ions are coordinated in B

DOI: 10.1021/acs.cgd.9b00880 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Communication

100 to 350 K, although the principal axis X2 shows relatively small PTE when the PASCal calculation was done from 100 to 450 K (see Table S11 and Figure S7 in Supporting Information). The expansivity indicatrix and percentage change in principal axes are shown in Figure 3a,b, respectively.

octahedral geometry with three oxygen atoms from three carboxylate ions of H3BTC, a nitrogen atom of the bpy ligand, and two OH− ions. Both metal ions along with carboxylate and hydroxide ions construct a cluster (Figure 1b). The cluster serves as a secondary building unit (SBU) which are linked with each other through the HBTC2− ion resulting in a linear architecture propagating parallel to the crystallographic a axis (Figure 1c). Linear arrays are bridged by the symmetric bpy ligands coordinated with Co(II) leading to the formation of 2D layers parallel to the (0 1 0) planes (Figure 1d). 2D layers are connected by O−H···O and O−H···N hydrogen bonding between carboxylic groups and the uncoordinated nitrogen atom of the asymmetric bpy molecules (Figure 1d). 2D layers are interlocked along the b axis by O−H···O hydrogen bonding. Interlocking of 2D layers generates a zip-like structure viewed down [1 0 1] (Figure 2).

Figure 2. Interlocking of stacked 2D layers connected by O−H···O hydrogen bonding forms a zip-like structure.

VT-SCXRD experiments were performed in order to understand the thermal expansion property of 1. Due to limitations in the thermostat, diffraction experiments were carried out until 450 K, although the material is thermally stable until around 600 K as observed in the thermogravimetric analysis (TGA) (see Figure S2 in Supporting Information). Diffraction data were recorded at each temperature at intervals of 50 K starting at 100 to 450 K, and crystal structures of 1 at each temperature have been solved and refined. (For crystallographic parameters and refinement details, see Table S1 in Supporting Information.) Initial analysis of unit cell parameters reveals the usual expansion of unit cell axes with increasing temperature, whereas the interaxial angles α and β increase and γ decreases with increasing temperature. Relatively large expansion of the b axis in comparison to the other two axes prompted us for a detailed investigation of the thermal expansion property of 1. Since the crystal system is nonorthogonal, thermal expansion coefficients were calculated using the PASCal program.59 The coefficient of thermal expansion (CTE) of the principal axes X1, X2, and X3 within the temperature range of 100 to 350 K are −36(2), +2(1), and +152(3) MK−1 along the approximate direction of [−3 2 −1], [0 1 1], and [−1 −2 1], respectively (see Table S10 and Figure S6 in Supporting Information). In other words, principal axes X1, X2, and X3 exhibit NTE, near ZTE, and PTE, respectively, within the temperature range of

Figure 3. (a) Expansivity indicatrix for temperature range from 100 to 350 K obtained from the PASCal program.57 (b) Percentage change in length with temperature from 100 to 350 K along X1, X2, and X3 principal axes obtained from the PASCal program.59

The reversibility of the thermal expansion property of 1 has been verified by determination of unit cell parameters using SCXRD initially at 100 K, then heated to 450 K, and again cooling back to 100 K (see Table S2 in Supporting Information). Detailed analysis of the crystal structures determined at each temperature reveals the origin of axial NTE, PTE, and near ZTE in 1. A significant increase of some bond angles between metal ions, carboxyl groups, and hydroxyl group has been observed with increasing temperature. For example, ∠C(16) O(7) Zn(1) increases from 111.7(1)° to 113.1(2)°; ∠C(17) O(10) Co(1) changes from 121.7(1)° to 124.0(2)°; ∠C(16) O(8) Co(1) increases from 143.0(1)° to 144.5(2)°, and ∠Co(1) O(13) Zn(1) increases from 121.4(1)° to 122.7(1)° (see Table S7). The angular change leads to the change of intermetallic distance in 2D layers resulting in a systematic change in the crystal structure. The 2D layer has been shown schematically in Figure 4a. In order to explain the change of C

DOI: 10.1021/acs.cgd.9b00880 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Communication

intermetallic distances, a rectangular block between four SBU has been chosen (Figure 4b).

Figure 5. Variation of metal−metal distance with temperature in 2D layer denoted by ui, vi, and wi as shown in Figure 4a.

Analysis of crystal structures also revealed that O−O distances of O−H···O hydrogen bonding between 2D layers in the crystal structure remain almost unaffected with increasing temperature (Table S3 in Supporting Information). Relatively large changes observed in the distances ui, vi, and wi above 350 K are possibly due to the high thermal vibrational motion of the atoms. In summary, a new mixed-metal mixed-linker 2D coordination framework material synthesized by hydrothermal reaction exhibits axial NTE, PTE, and ZTE in the temperature range from 100 to 350 K. PTE and NTE in compound 1 originated in the 2D layer of the crystal structure due to flexible coordination of organic ligands with Co(II) and Zn(II) ions in octahedral and tetrahedral geometry, respectively. On the other hand, thickness of the 2D layers and interlayer distances have not been changed with the change of temperature leading to ZTE. Understanding of crystal packing in this study could be useful to designing new framework materials using mixed metals and mixed ligands to tune thermal expansion properties that could be useful for various applications in thermomechanical sensors and actuators.

Figure 4. (a) Schematic diagram of a 2D layer in the crystal structure of 1. Only SBUs in the layers have been shown and connecting ligands are omitted for clarity. Part of the organic ligands connecting SBUs have been shown by bars. Co(II) and Zn(II) ions are shown in purple and cyan. (b) Schematic representation of a block in the 2D layer shown in Figure 4a to describe the change in the metal−metal distances represented by ui, vi, and wi along the principal axes X2, X3, and X1, respectively.

The points A, B, C, and D indicate the position of Zn(II) ions of SBUs which are immediate neighbors to each other and related by translational symmetry. Likewise, the points E, F, G, and H are the positions of translationally related Co atoms of the same SBUs. Thickness of the 2D layer (i.e., AE = BF = CG = DH), diagonal distances (i.e., BD = FH and AC = EG) are designated as ui, vi, and wi (i = 1 or 2), respectively. Thermal expansion of 1 along principal axes X1, X2, and X3 can be rationalized based on variation of ui, vi, and wi with the change of temperature (Figure 5 and Table S6 in Supporting Information). The diagonal distance v1 (16.653(1) Å) increases to v2 (16.872(1) Å) as the temperature increases from 100 to 350 K. At the same time, the distance w1 (18.578(1) Å) deceases to w2 18.527(1) Å. Expansion of vi and contraction of wi of the layer has been observed along the approximate direction of the principal axes X3 and X1, respectively, along which PTE and NTE occur. On the other hand, thickness of the layer, ui shown in Figure 4b changes from 3.1139(5) Å to 3.1163(5) Å, which implies that ui remains almost constant as temperature increases from 100 to 350 K. The direction of ui is almost along the approximate direction of the principal axis X2 along which ZTE occurs. Large expansion of the crystallographic b axis in the crystal structure of compound 1 can be attributed to the angular change of the coordination angle around the metal ions, which is consequential in the large thermal expansion along the principal axis X3 (Table S7 in Supporting Information).



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.cgd.9b00880. Synthesis, VT-SCXRD, Thermal ellipsoid plots, PXRD patterns, TGA plots, Hydrogen bonding parameters, Crystallographic details, Thermal expansion coefficient (PDF) Accession Codes

CCDC 1895733−1895740 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], or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033. D

DOI: 10.1021/acs.cgd.9b00880 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design



Communication

(12) 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. (13) Cai, W.; Katrusiak, A. Giant Negative Linear Compression Positively Coupled to Massive Thermal Expansion in a Metal-Organic Framework. Nat. Commun. 2014, 5, 4337. (14) 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. (15) Chen, J.; Hu, L.; Deng, J.; Xing, X. Negative Thermal Expansion in Functional Materials: Controllable Thermal Expansion by Chemical Modifications. Chem. Soc. Rev. 2015, 44, 3522−3567. (16) Shrivastava, A.; Negi, L.; Das, D. Area Negative Thermal Expansion in a Mixed Metal Mixed Organic MOF: “ElevatorPlatform” Mechanism Induced by O-H···O Hydrogen Bonding. CrystEngComm 2018, 20, 4719−4723. (17) Birkedal, H.; Schwarzenbach, D.; Pattison, P. Observation of Uniaxial Negative Thermal Expansion in an Organic Crystal. Angew. Chem., Int. Ed. 2002, 41, 754−756. (18) Haas, S.; Batlogg, B.; Besnard, C.; Schiltz, M.; Kloc, C.; Siegrist, T. Large Uniaxial Negative Thermal Expansion in Pentacene Due to Steric Hindrance. Phys. Rev. B: Condens. Matter Mater. Phys. 2007, 76, 205203. (19) Das, D.; Jacobs, T.; Barbour, L. J. Exceptionally Large Positive and Negative Anisotropic Thermal Expansion of an Organic Crystalline Material. Nat. Mater. 2010, 9, 36−39. (20) Fortes, A. D.; Suard, E.; Knight, K. S. Negative Linear Compressibility and Massive Anisotropic Thermal Expansion in Methanol Monohydrate. Science 2011, 331, 742−746. (21) Saha, B. K. Thermal Expansion in Organic Crystals. J. Indian Inst. Sci. 2017, 97, 177−191. (22) Das, D.; Barbour, L. J. Uniaxial Negative Thermal Expansion Induced by Moiety Twisting in an Organic Crystal. CrystEngComm 2018, 20, 5123−5126. (23) van der Lee, A.; Roche, G. H.; Wantz, G.; Moreau, J. J. E.; Dautel, O. J.; Filhol, J.-S. Experimental and Theoretical Evidence of a Supercritical-like Transition in an Organic Semiconductor Presenting Colossal Uniaxial Negative Thermal Expansion. Chem. Sci. 2018, 9, 3948−3956. (24) Negi, L.; Shrivastava, A.; Das, D. Switching from Positive to Negative Axial Thermal Expansion in Two Organic Crystalline Compounds with Similar Packing. Chem. Commun. 2018, 54, 10675− 10678. (25) Salvador, J. R.; Guo, F.; Hogan, T.; Kanatzidis, M. G. Erratum: Zero Thermal Expansion in YbGaGe Due to an Electronic Valence Transition. Nature 2003, 425, 702−705. (26) Margadonna, S.; Prassides, K.; Fitch, A. N. Zero Thermal Expansion in a Prussian Blue Analogue. J. Am. Chem. Soc. 2004, 126, 15390−15391. (27) Kelly, A.; McCartney, L. N.; Clegg, W. J.; Stearn, R. J. Controlling Thermal Expansion to Obtain Negative Expansivity Using Laminated Composites. Compos. Sci. Technol. 2005, 65, 47−59. (28) Liu, Q. Q.; Cheng, X. N.; Yang, J. Development of low thermal expansion Sc2(WO4)3 containing composites. Mater. Technol. 2012, 27, 388−392. (29) Evans, J. S. O. Negative Thermal Expansion Materials. J. Chem. Soc., Dalton Trans. 1999, 19, 3317−3326. (30) Takenaka, K. Negative Thermal Expansion Materials: Technological Key for Control of Thermal Expansion. Sci. Technol. Adv. Mater. 2012, 13, 013001. (31) 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. (32) Chapman, K. W.; Chupas, P. J.; Kepert, C. J. Direct Observation of a Transverse Vibrational Mechanism for Negative Thermal Expansion in Zn(CN)2: An Atomic Pair Distribution Function Analysis. J. Am. Chem. Soc. 2005, 127, 15630−15636.

AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Tel: +91-2673-8813. ORCID

Dinabandhu Das: 0000-0003-0143-7821 Funding

This work was supported by DST-SERB (Project of Id: SB/ FT/CS-064/2014) and PURSE project in JNU sponsored by DST, India. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS AS is grateful to Science and Engineering Research Board (SERB), DST, India for financial support as JRF. AS acknowledges Lalita Negi for assisting in diffraction experiment. DD thanks to SERB, DST, India, for the project of Id: SB/FT/CS-064/2014 and DST for PURSE project in JNU. We are grateful to DST for providing Single Crystal XRD and TGA facilities in SPS, JNU under FIST program.



REFERENCES

(1) Panda, M. K.; Runčevski, T.; Sahoo, S. C.; Belik, A. A.; Nath, N. K.; Dinnebier, R. E.; Naumov, P. Colossal Positive and Negative Thermal Expansion and Thermosalient Effect in a Pentamorphic Organometallic Martensite. Nat. Commun. 2014, 5, 4811−1−4811−8. (2) 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. (3) 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, 794−798. (4) Korčok, J. L.; Katz, M. J.; Leznoff, D. B. Impact of Metallophilicity on “Colossal” Positive and Negative Thermal Expansion in a Series of Isostructural Dicyanometallate Coordination Polymers. J. Am. Chem. Soc. 2009, 131, 4866−4871. (5) 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. (6) 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. (7) Adak, S.; Daemen, L. L.; Hartl, M.; Williams, D.; Summerhill, J.; Nakotte, H. Thermal Expansion in 3d-Metal Prussian Blue Analogs A Survey Study. J. Solid State Chem. 2011, 184, 2854−2861. (8) Gao, Q.; Shi, N.; Sanson, A.; Sun, Y.; Milazzo, R.; Olivi, L.; Zhu, H.; Lapidus, S. H.; Zheng, L.; Chen, J.; Xing, X. Tunable Thermal Expansion from Negative, Zero, to Positive in Cubic Prussian Blue Analogues of GaFe(CN)6. Inorg. Chem. 2018, 57, 14027−14030. (9) 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. (10) Devries, L. D.; Barron, P. M.; Hurley, E. P.; Hu, C.; Choe, W. Nanoscale Lattice Fence” in a Metal-Organic Framework: Interplay between Hinged Topology and Highly Anisotropic Thermal Response. J. Am. Chem. Soc. 2011, 133, 14848−14851. (11) Azuma, M.; Chen, W. T.; Seki, H.; Czapski, M.; Olga, S.; Oka, K.; Mizumaki, M.; Watanuki, T.; Ishimatsu, N.; Kawamura, N.; Ishiwata, S.; Tucker, M. G.; Shirakawa, Y.; Attfield, J. P. Colossal Negative Thermal Expansion in BiNiO3 Induced by Intermetallic Charge Transfer. Nat. Commun. 2011, 2, 345−347. E

DOI: 10.1021/acs.cgd.9b00880 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Communication

(33) Wu, Y.; Peterson, V. K.; Luks, E.; Darwish, T. A.; Kepert, C. J. Interpenetration as a Mechanism for Negative Thermal Expansion in the Metal-Organic Framework Cu3(Btb)2 (MOF-14). Angew. Chem., Int. Ed. 2014, 53, 5175−5178. (34) Li, B.; Luo, X. H.; Wang, H.; Ren, W. J.; Yano, S.; Wang, C.-W.; Gardner, J. S.; Liss, K.-D.; Miao, P.; Lee, S.-H.; Kamiyama, T.; Wu, R. Q.; Kawakita, Y.; Zhang, Z. D. Colossal Negative Thermal Expansion Induced by Magnetic Phase Competition on Frustrated Lattices in Laves Phase Compound (Hf,Ta) Fe2. Phys. Rev. B: Condens. Matter Mater. Phys. 2016, 93, 3−8. (35) 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. (36) Takenaka, K. Progress of Research in Negative Thermal Expansion Materials: Paradigm Shift in the Control of Thermal Expansion. Front. Chem. 2018, 6, 267. (37) Arvanitidis, J.; Papagelis, K.; Margadonna, S.; Prassides, K.; Fitch, A. N. Temperature-Induced Valence Transition and Associated Lattice Collapse in Samarium Fulleride. Nature 2003, 425, 599−602. (38) Welche, P. R. L.; Heine, V.; Dove, M. T. Negative Thermal Expansion in Beta-Quartz. Phys. Chem. Miner. 1998, 26, 63−77. (39) Collings, I. E.; Tucker, M. G.; Keen, D. A.; Goodwin, A. L. Geometric Switching of Linear to Area Negative Thermal Expansion in Uniaxial Metal-Organic Frameworks. CrystEngComm 2014, 16, 3498−3506. (40) Hemberger, J.; Nidda, H.-A. K. v.; Tsurkan, V.; Loidl, A. Large Magnetostriction and Negative Thermal Expansion in the Frustrated Antiferromagnet ZnCr2Se4. Phys. Rev. Lett. 2007, 98, 1−4. (41) 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. (42) Ferey, G. Hybrid porous solids: past, present, future. Chem. Soc. Rev. 2008, 37, 191−214. (43) Murray, L. J.; Dinca, M.; Long, J. R. Hydrogen storage in metal−organic frameworks. Chem. Soc. Rev. 2009, 38, 1294−1314. (44) Zhang, W.; Xiong, R.-G. Ferroelectric Metal−Organic Frameworks. Chem. Rev. 2012, 112, 1163−1195. (45) Sumida, K.; Rogow, D. L.; Mason, J. A.; McDonald, T. M.; Bloch, E. D.; Herm, Z. R.; Bae, T.-H.; Long, J. R. Carbon Dioxide Capture in Metal−Organic Frameworks. Chem. Rev. 2012, 112, 724− 781. (46) Talin, A. A.; Centrone, A.; Ford, A. C.; Foster, M. E.; Stavila, V.; Haney, P.; Kinney, R. A.; Szalai, V.; Gabaly, F. E.; Yoon, H. P.; Leonard, F.; Allendorf, M. D. Tunable Electrical Conductivity in Metal-Organic Framework Thin-Film Devices. Science 2014, 343, 66− 70. (47) Furukawa, H.; Cordova, K. E.; O’Keeffe, M.; Yaghi, O. M. The Chemistry and Applications of Metal-Organic Frameworks. Science 2013, 341, 1230444. (48) Lee, J. Y.; Farha, O. K.; Roberts, J.; Scheidt, K. A.; Nguyen, S. T.; Hupp, J. T. Metal−organic framework materials as catalysts. Chem. Soc. Rev. 2009, 38, 1450−1459. (49) Wang, Z.; Knebel, A.; Grosjean, S.; Wagner, D.; Bräse, S.; Wöll, C.; Caro, J.; Heinke, L. Tunable Molecular Separation by Nanoporous Membranes. Nat. Commun. 2016, 7, 13872. (50) Stassen, I.; Burtch, N.; Talin, A.; Falcaro, P.; Allendorf, M.; Ameloot, R. An Updated Roadmap for the Integration of MetalOrganic Frameworks with Electronic Devices and Chemical Sensors. Chem. Soc. Rev. 2017, 46, 3185−3241. (51) Wu, M. X.; Yang, Y. W. Metal−Organic Framework (MOF)Based Drug/Cargo Delivery and Cancer Therapy. Adv. Mater. 2017, 29, 1−20. (52) Lu, W.; Wei, Z.; Gu, Z. Y.; Liu, T. F.; Park, J.; Park, J.; Tian, J.; Zhang, M.; Zhang, Q.; Gentle, T.; Bosch, M.; Zhou, H.-C. Tuning the Structure and Function of Metal-Organic Frameworks via Linker Design. Chem. Soc. Rev. 2014, 43, 5561−5593. (53) Wang, H.; Dong, X.; Lin, J.; Teat, S. J.; Jensen, S.; Cure, J.; Alexandrov, E. V.; Xia, Q.; Tan, K.; Wang, Q.; Olson, D. H.;

Proserpio, D. M.; Chabal, Y. J.; Thonhauser, T.; Sun, J.; Han, Y.; Li, J. Topologically Guided Tuning of Zr-MOF Pore Structures for Highly Selective Separation of C6 Alkane Isomers. Nat. Commun. 2018, 9, 1745. (54) Yang, C.; Wang, X.; Omary, M. A. Crystallographic Observation of Dynamic Gas Adsorption Sites and Thermal Expansion in a Breathable Fluorous Metal-Organic Framework. Angew. Chem., Int. Ed. 2009, 48, 2500−2505. (55) Zhou, H.-L.; Zhang, Y.-B.; Zhang, J.-P.; Chen, X.-M. Supramolecular-jack-like guest in ultramicroporous crystal for exceptional thermal expansion behaviour. Nat. Commun. 2015, 6, 6917. (56) Zhou, H.-L.; Zhang, J.-P.; Chen, X.-M. Controlling Thermal Expansion Behaviors of Fence-Like Metal-Organic Frameworks by Varying/Mixing Metal Ions. Front. Chem. 2018, 6, 306. (57) Rodrigues, A. D.; Fahsi, K.; Dumail, X.; Masquelez, N.; Lee, A. v. d.; Mallet-Ladeira, S.; Sibille, R.; Filhol, J.-S.; Dutremez, S. G. Joint Experimental and Computational Investigation of the Flexibility of a Diacetylene-Based Mixed-Linker MOF: Revealing the Existence of Two Low-Temperature Phase Transitions and the Presence of Colossal Positive and Giant Negative Thermal Expansions. Chem. Eur. J. 2018, 24, 1586−1605. (58) Lama, P.; Das, R. K.; Smith, V. J.; Barbour, L. J. A Combined Stretching-Tilting Mechanism Produces Negative, Zero and Positive Linear Thermal Expansion in a Semi-Flexible Cd(II)-MOF. Chem. Commun. 2014, 50, 6464−6467. (59) Cliffe, M. J.; Goodwin, A. L. PASCal: A Principal-Axis Strain Calculator for Thermal Expansion and Compressibility Determination. J. Appl. Crystallogr. 2012, 45, 1321−1329.

F

DOI: 10.1021/acs.cgd.9b00880 Cryst. Growth Des. XXXX, XXX, XXX−XXX