Low-Frequency Phonon Driven Negative Thermal Expansion in Cubic

Aug 14, 2018 - Beijing Advanced Innovation Center for Materials Genome Engineering .... Zalibera, Bučinský, Malček, Govind, Karunakaran, Farid, McD...
0 downloads 0 Views 2MB Size
Article Cite This: Inorg. Chem. XXXX, XXX, XXX−XXX

pubs.acs.org/IC

Low-Frequency Phonon Driven Negative Thermal Expansion in Cubic GaFe(CN)6 Prussian Blue Analogues Qilong Gao,†,‡ Naike Shi,†,‡ Qiang Sun,§ Andrea Sanson,∥ Ruggero Milazzo,∥ Alberto Carnera,∥ He Zhu,†,‡ Saul H. Lapidus,⊥ Yang Ren,⊥ Qingzhen Huang,# Jun Chen,*,†,‡ and Xianran Xing†,‡

Downloaded via KAOHSIUNG MEDICAL UNIV on August 15, 2018 at 07:16:00 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.



Beijing Advanced Innovation Center for Materials Genome Engineering, University of Science and Technology Beijing, Beijing 100083, China ‡ Department of Physical Chemistry, University of Science and Technology Beijing, Beijing 100083, China § International Laboratory for Quantum Functional Materials of Henan, School of Physics and Engineering, Zhengzhou University, Zhengzhou 450001, China ∥ Department of Physics and Astronomy, University of Padova, Padova I-35131, Italy ⊥ Argonne National Laboratory, X-ray Science Division, Argonne, Illinois 60439, United States # NIST Center for Neutron Research, National Institute of Standards and Technology, Gaithersburg, Maryland 20899-6102, United States S Supporting Information *

ABSTRACT: The understanding of the negative thermal expansion (NTE) mechanism is vital not only for the development of new NTE compounds but also for effectively controlling thermal expansion. Here, we report an interesting isotropic NTE property in cubic GaFe(CN)6 Prussian blue analogues (αl = −3.95 × 10−6 K−1, 100−475 K), which is a new example to understand the complex NTE mechanism. A combined study of synchrotron X-ray diffraction, X-ray total scattering, X-ray absorption fine structure, neutron powder diffraction, and density functional theory calculations shows that the NTE of GaFe(CN)6 originates from the lowfrequency phonons (< ∼100 cm−1), which are directly related to the transverse vibrations of the atomic −Ga−NC−Fe− chains. Both the Ga−N and Fe−C chemical bonds are much softer to bend than to stretch. The direct evidence that transverse vibrational contribution to the NTE of GaFe(CN)6 is dominated by N, instead of C atoms, is illustrated. It is interesting to find that the polyhedra of GaFe(CN)6 are not rigid, which is a starting assumption in some models describing the NTE properties of other systems. The NTE mechanism can be vividly described by the “guitar-string” effect, which would be the common feature for the NTE property of many open-framework functional materials, such as Prussian blue analogues, oxides, cyanides, metal− organic frameworks, and zeolites.



INTRODUCTION Negative thermal expansion (NTE) is an extraordinary phenomenon, which attracts considerable attention not only for the understanding of the nature of thermal expansion of materials but also for potential applications needing controlling of thermal expansion.1−7 Great achievements have been obtained in the last two decades, in which a large number of NTE materials have been discovered and various mechanisms have been disclosed for the NTE phenomena. In general, NTE compounds can be classified into two main categories: one is driven by low-frequency phonons in the framework, such as oxides,1,8−11 fluorides,12−14 cyanides,15−21 and MOFs;22−25 the other is driven by changing in electronic state, such as magnetovolume effect in magnetics,2,26 ferroelectrovolume effect in ferroelectrics,4,27,28 intermetallic charge transfer,29,30 and Mott phase transition like in Ca2RuO4.31,32 Since the © XXXX American Chemical Society

discovery of ZrW2O8 in 1996, the rigid unit modes (RUMs) were widely employed in order to explain the origin of NTE in open-framework structures. Understanding of the NTE mechanism is one of the critical issues for the development of these materials. However, there are still controversies. For example, the nature of the NTE prototype compound ZrW2O8 remains debated at the present. According to the “RUMs”, the NTE of ZrW2O8 is a result from the fact that transverse vibrations of the oxygen atoms in the Zr-O-W linkage lead to the rocking motion of ZrO6 and WO4 rigid polyhedra.33 Nevertheless, the “tent model” suggested that NTE is derived from translational motion of WO4 tetrahedra along the ⟨111⟩ axis and the correlated motion of the three nearest ZrO6 Received: June 5, 2018

A

DOI: 10.1021/acs.inorgchem.8b01526 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 1. (a) High-resolution SXRD refinement of GaFe(CN)6 at 300 K. The inset shows the framework structure, consisting of FeC6 and GaN6 octahedra linked by CN ligands. (b) Temperature dependence of lattice constant derived from both high-resolution SXRD and NPD.

octahedra.34 Molecular dynamics study showed that a more flexible model, simply based on rigid nearest W−O and Zr−O bonds and tension effect, should be adopted.35 More recently, the study on the NTE mechanism of ScF3 revealed that the transverse vibrations of fluorine atoms are critical for NTE.14 It has been well-known that Prussian blue analogues (PBAs) exhibit intriguing physical or chemical properties such as magnetism,36 energy storage,37 environmental catalysis,38 and drug delivering.39 In 2004, the FeCo(CN)6 PBAs were reported to exhibit zero thermal expansion property.17 Interesting NTE properties have been later discovered in various compounds, such as Zn(CN)2,15 ZnPt(CN)6,18 Ag3Co(CN)6,19 LnCo(CN)6,20 and YFe(CN)6.21 The PBAs have a unique structure of chemical bonds, where the flexible linkages of -M-CN-M- (M is metal) consist of an openframework structure. As a member of NTE open-framework materials, the family of cyanides and PBAs is a good candidate for exploring new NTE materials and understanding the complicated NTE mechanism. Here, an interesting isotropic NTE has been found in the cubic GaFe(CN)6 of PBAs. The study of its local structure and vibrational dynamics reveals deeper insight in the NTE mechanism. Accordingly, a combined study of high-resolution synchrotron X-ray diffraction (SXRD), neutron powder diffraction (NPD), pair distribution function (PDF), extended X-ray absorption fine structure (EXAFS), and first-principles calculations has been performed. The key factor for the NTE behavior of GaFe(CN)6 is the presence of large transverse vibrations from the cyanide bridges. The NTE mechanism of the present open-framework structure PBAs interpreted within the “guitar-string” effect, i.e., the contribution to the NTE of lattices from the tension effect that overwhelms the one to the PTE of chemical bonds from the bond-stretching.



NIST Center for Neutron Research on the BT-1 high-resolution neutron powder diffractometer (λ = 1.5397 Å). The crystal structure was resolved by the Rietveld method, using the FULLPROF program.40 The pair distribution functions were collected at the beamline 11-ID-C of APS, Argonne National Laboratory, with X-ray wavelength, λ, of 0.1173 Å. The G(r) functions were obtained by PDFgetX2 and G(r) = 4πr[ρ(r) − ρ0], where ρ(r) and ρ0 are the instantaneous and average densities.41 In order to compare different temperature peak positions, the G(r) curves were converted to a radial distribution function R(r), [R(r) = r(G(r) + 4πrρ0].42 The morphology of the sample was examined using a field-emission scanning electron microscope (FE-SEM, SUPRA-40, Carl Zeiss). Thermal stability of GaFe(CN)6·2H2O was investigated in N2 condition with a heating rate of 1 K/min by TG/DSC analysis (Labsys Evo, Setaram Corp.). EXAFS Measurements. Ga and Fe K-edge EXAFS measurements of GaFe(CN)6 were performed from 475 to 300 K with a step of 35 K at the XAFS beamline of ELETTRA synchrotron in Trieste (Italy). The sample for EXAFS was prepared by mixing and pelletizing the sample powder with boron nitride. The amount of sample powder was chosen so as to have an absorption edge jump Δμx ∼ 1. The EXAFS spectra were collected in transmission mode in the energy range of ∼10.1−11.6 for Ga and 6.8−8.4 keV for Fe, with an energy step varying from 0.2 eV in the near-edge region to about 4.5 eV at the highest energies, thus to obtain a uniform wave vector step Δk ∼ 0.035 Å−1. The X-ray beam was monochromatized by a Si(111) double-crystal monochromator. The samples, kept under highvacuum (10−5 mbar) during the entire experiment, was mounted in a high-temperature furnace, and the temperature was stabilized and monitored through an electric heater controlled by a feedback loop, ensuring a thermal stability within ±1 K. Two spectra were collected at each temperature point. The EXAFS signals, extracted according to well-established procedures,43,44 are shown in Figure S1. In order to separate the contributions of the different coordination shells, the k-weighted EXAFS signals were Fourier transformed (FT) in the k range 2−14 Å−1 using a Gaussian window (Figure S2). The FT structure between about 0.5 and 3 Å is mainly due to the six nearest-neighbor N atoms around Ga and six nearest-neighbor C atoms around Fe. The other scattering paths contributing to the FT peak up to about 3 Å have been calculated using the FEFF code45 and are listed in Tables S1 and S2. A nonlinear best fit to the experimental spectra was then performed in the r-space between 0.5 and 3.0 Å (bold-dashed lines in Figure S3) by using the FEFFIT package.46 The vibrational dynamics parameters of the Ga−N and Fe−C atomic pairs are reported in Table S3. Computational Methods. The first-principles density functional theory (DFT) calculations were performed using the Vienna Ab initio Simulation Package (VASP)47 with the projector augmented wave (PAW) method.48 For the treatment of exchange-correlation energy, we employed the generalized gradient approximation (GGA) functional of Perdew−Burke−Ernzerhof (PBE).49 The kinetic-energy

EXPERIMENTAL SECTION

Materials Synthesis. The powder sample of GaFe(CN)6·2H2O was prepared by a solution precipitation method. 50 mL of 0.1 moL L−1 Ga(NO3)3 was added to 50 mL of 0.1 moL L−1 K3Fe(CN)6 aqueous solution. The mixture solution was maintained for 10 h at 60 °C under vigorous stirring. The blue green precipitation was collected by filtration, washed many times with water and ethanol, and then dried at 50 °C for 10 h. Finally, the sample was kept in a black screw cap vial. The anhydrous sample of GaFe(CN)6 was obtained after the dehydration of GaFe(CN)6·2H2O by heating at 202 °C for 10 h. Sample Characterizations. Crystal structure was characterized by high-resolution synchrotron XRD (SXRD), which was performed at the 11-BM-B beamline of Advanced Photon Source (λ = 0.412634 Å). Neutron powder diffraction (NPD) data were collected at the B

DOI: 10.1021/acs.inorgchem.8b01526 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 2. (a) Schematic diagram of Ga−NC−Fe ligands, where the transverse vibration of CN is represented. Temperature dependence of the peak position for the (b) Ga−N/Fe−C, and (c) Ga···Fe atomic pairs. (d) Comparison between the “apparent” and the “true” Ga−N and Fe−C bond lengths in GaFe(CN)6 as a function of temperature. The “apparent” bond lengths are measured by high-resolution synchrotron XRD, while the “true” ones by PDF and EXAFS. cutoff of the plane-wave basis set was taken to be 600 eV, and the kspace integration was performed with the Monkhorst−Pack meshes of (3 × 3 × 3). Convergence criteria for the total energy and the ionic relaxation were 10−8 eV and 10−4 eV/Å, respectively. Vibrational properties were calculated through the PHONOPY code,50 where the real space force constants were calculated by employing a 10.19 Å × 10.19 Å × 10.19 Å unit cell. The mode Grüneisen parameter is calculated by γi = −

∂ ln ωi ∂ ln V

metal molecular poly-cyanide open framework (Figure 1a), which consists of a 3D cubic network with a linear bridging of alternating FeC6 and GaN6 octahedra. We point out that the crystal structure of the present GaFe(CN)6 is an ideal example to study the NTE behavior. Compared with other NTE PBAs like Zn3[Fe(CN)6]251 and Rb0.97Mn[Fe(CN)6]0.99·0.3H2O,52 there are no vacancies or cationic guests in the structure of GaFe(CN)6. It needs to be mentioned that other research of Mössbauer spectra and electrochemical behavior have been reported in hydrated GaFe(CN)6.53−56 The lattice constant of GaFe(CN)6 was determined from temperature dependence of both high-resolution SXRD and NPD. It is interesting to notice that GaFe(CN)6 exhibits an isotropic NTE property, where the lattice constant contracts almost linearly with increasing temperature (Figure 1b). The average linear thermal expansion coefficient (αl) is −3.95(5) × 10−6 K−1 (100−475 K). Compared with other cubic NTE compounds, the present GaFe(CN)6 shows a moderate NTE, stronger than FeCo(CN)617 (−1.47 × 10−6 K−1, 4.2−300 K) or NiPt(CN)6 (αl = −1.02 × 10−6 K−1, 100−330 K),16 but weaker than Cd(CN)2 (−20.4 × 10−6 K−1, 150−375 K),15 CdPt(CN)6 (−10.02 × 10−6 K−1, 100−240 K),18 and ZrW2O8 (−9.1 × 10−6 K−1, 0.4−430 K).15 Since GaFe(CN)6 has a cubic structure, the chemical bonds can be calculated according to the structural refinement results. However, we recall that diffraction methods give the “apparent” bond length (i.e., distance between the atomic average positions), but not the “true” bond length (i.e., average of the atomic pair distance). Accordingly, we adopt here local structure methods

(1)

where ωi is the frequency of the ith mode, and V is the volume of the unit cell.



RESULTS AND DISCUSSION The sample of GaFe(CN)6·2H2O was prepared by the reaction of aqueous solutions of Ga(NO3)3 and K3Fe(CN)6 at 60 °C under vigorous stirring for 10 h. The as-prepared sample of GaFe(CN)6·2H2O exhibits a cube shape with the size of 100− 300 nm (see Figure S4 in the Supporting Information). As shown in the TG−DSC result (see Figure S5 in the Supporting Information), two water molecules in the framework may be considered as uncoordinated or zeolitic guest water molecules. After the dehydration at 202 °C, the sample of GaFe(CN)6 was obtained. Here, the structural properties of GaFe(CN)6·2H2O and GaFe(CN)6 were studied by high-resolution SXRD and NPD (Figure 1a; see Figures S6−S8 and Tables S4 and S5 in the Supporting Information). The space group results Fm3̅m. Similarly to most PBAs, GaFe(CN)6 has a flexible doubleC

DOI: 10.1021/acs.inorgchem.8b01526 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 3. Temperature dependence of (a) ADPs of the N and C atoms, (b) Ga-N and (c) Fe-C perpendicular (⊥) and parallel (∥) MSRDs, and (d) anisotropy of the relative thermal vibrations of the Ga−N and Fe−C atomic pairs of GaFe(CN)6.

as PDF and EXAFS to study the behavior of the “true” bond lengths. Figure 2a shows the schematic of CN ligands dynamic vibration, where the difference between “apparent” and “true” bond lengths can be distinguished. The “apparent” bond length is obtained from the distance between the average positions of two atoms as extracted from the high-resolution SXRD measurements, while the local structure PDF and EXAFS analyses were performed for the “true” bond lengths. Here, the change of “true” bond lengths can be revealed through the peak position in the G(r) as a function of temperature. It is interesting to observe that the pair distance of the Ga−N and Fe−C chemical bonds actually expands with increasing temperature (Figure 2b), thus disclosing an intrinsic positive thermal expansion (PTE) property. We have to remark that such pair distance is the average value from both Ga−N and Fe−C bonds. Attempts to any further disentanglement between the Ga−N and Fe−C bonds are unreliable, because the difference among them is less than 0.1 Å. As shown in Figure 2c, the direct experimental evidence for the NTE property of GaFe(CN)6 can be observed, according to the shrinkage in the atomic pair of the Ga···Fe linkages with increasing temperature. It is in good agreement with the lattice constant behavior (Figure 1b). In order to distinguish the Ga− N and Fe−C bonds, temperature-dependent Ga and Fe K-edge EXAFS measurements have been performed. As shown in Figure 2d, an opposite trend of the chemical bonds of Ga−N and Fe−C can be observed by comparing the high-resolution SXRD and EXAFS. On heating, the “apparent” bond length measured by SXRD strongly contracts (αl = −2.30 × 10−5 K−1 for Ga−N, αl = −0.56 × 10−5 K−1 for Fe−C), while the “true”

one, measured by EXAFS, oppositely expands (αl = +3.76 × 10−5 K−1, αl = +2.28 × 10−5 K−1 for Fe−C). A similar behavior was also observed in other NTE compounds like ZrW2O8,35 Zn(CN)2,16 and ScF3.14 PTE might be the common behavior for NTE materials, which originates from the nature of bondstretching effect. These results on the “apparent” and “true” bond length suggest that the NTE is strictly related to a specific lattice dynamics of −M−CN−M− linkages. In order to deeply understand the NTE mechanism of GaFe(CN)6, the lattice dynamics of N and C atoms have been studied by both anisotropic atomic displacement parameters (ADPs) and atomic mean square relative displacements (MSRDs). The ADPs, which can be reliably extracted from NPD because of pronounced different neutron scattering lengths of C (0.6646 cm−12) and N (0.936 cm−12), do not include the correlation of atomic motion, which are taken into account for the MSRDs as measured by EXAFS instead.57 Since the ADPs of Fe and Ga are isotropic and their values are much smaller compared with C and N atoms, they do not affect the transverse thermal vibration of C and N atoms of GaFe(CN)6. As shown in Figure 3a, there is a pronouncedly anisotropic character in ADPs for both N and C atoms. The values of transverse ADPs (B22 = B33) are much larger than the longitudinal (B11) one, hence leading to a strong anisotropic lattice dynamics of both N and C atoms. However, the ADPs magnitude is much more intense in the N atoms, which suggests its main contribution to the transverse vibrations. Both the MSRDs parallel and perpendicular to the Ga−N and Fe−C bonds have been determined as a function of temperature (Figure 3b,c). They directly reveal that transverse relative vibrations of Ga−N and Fe−C pairs are generally D

DOI: 10.1021/acs.inorgchem.8b01526 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 4. (a) Phonon dispersion curves and phonon DOS. (b) The mode Grüneisen parameters as a function of frequency. The schematics of two representative transverse vibrational modes at the low frequencies of (c) 67 cm−1 and (d) 83.3 cm−1. The direction and size of arrows reflect the vibrational directions and amplitudes of the N and C atoms.

much larger than longitudinal. In particular, the ratio γ = MSRD⊥/MSRD∥ can be calculated to quantitatively describe the anisotropic vibrations of N and C atoms. Indeed, it is very large, about 51 ± 3 (Ga−N) and 24 ± 2 (Fe−C) at 475 K, (Figure 3d) so indicating that both Ga−N and Fe−C bonds are much softer to bend than to stretch. Large values of γ have also been observed in other NTE materials, such as Ag2O,58 ScF3,14 and ZrW2O8.35 By combining the results of PDF and EXAFS, it can be concluded that the NTE of GaFe(CN)6 comes from the strong tension effect, induced by the transverse vibrations of the atomic −M−CN−M− linkages. Moreover, the amplitude of transverse relative vibrations of the Ga−N bond is much larger than that of the Fe−C bond (Figure 3b,c), which also confirms that the most contribution of NTE is from N atoms. Finally, density functional theory (DFT) calculations have been carried out to further study the NTE mechanism of GaFe(CN)6. The crystallographic symmetry constraints from the experimental results were used. Only the phonon dispersion and the mode Grü neisen parameters were calculated. As shown in Figure 4a, in the low-frequency region below 100 cm−1, the phonon DOS of GaFe(CN)6 is strongly related to the vibrations of N atoms and less to the C atoms. Moreover, in this low-frequency region, most of the vibrational modes are with negative Grüneisen parameters (Figure 4b). This means that the transverse vibrations of N atoms represent the main contribution to the NTE of GaFe(CN)6, consistent with the results on both ADPs and MSRDs. Interestingly, the low-frequency mode at ∼67 cm−1, whose eigenvectors for the

C and N atoms have the same transverse direction (Figure 4c), possesses the largest negative Grüneisen parameter (−12.4). In this low-frequency mode, the vibration amplitude of N atoms is much larger than that of C atoms. As shown in Figure 4d, another interesting transverse vibration mode is noted at ∼83 cm−1, where C and N atoms have the opposite vibrational direction. However, such opposite transverse mode brings a relatively small contribution to NTE. The corresponding value of Grüneisen parameter is only about −6. The present calculations further indicate that the transverse vibrations of C and N atoms in the same direction contribute to the NTE of GaFe(CN)6 much more than those in the opposite way.



CONCLUSION According to these results on local structure and lattice dynamics, the thermal expansion occurs as the balance of two competitive roles: the tension effect to NTE and the bondstretching one to PTE. In the present work, the NTE behavior of GaFe(CN)6 is interpreted in terms of a prevalent tension effect to the bond-stretching one, which brings NTE in the “apparent” bond length and PTE in the “true” bond length, respectively. We propose a “guitar-string” effect to describe the occurrence of NTE in Prussian blue analogues of GaFe(CN)6. Here, the driving force for the occurrence of NTE is the transverse vibrations or rotations of C and N atoms, which act like the fingers of a guitar player. In summary, an interesting isotropic NTE (αl = −3.95(5) × 10−6 K−1, 100−475 K) has been found in cubic GaFe(CN)6. Even though its “true” bond lengths of Ga−N and Fe−C E

DOI: 10.1021/acs.inorgchem.8b01526 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry strongly expand with temperature, the “apparent” ones have an opposite NTE. This reflects the role of transverse vibrations of N and C atoms as driving force for the occurrence of NTE, especially for those of N atoms. Besides, the lattice dynamics calculation reveals that the low-frequency vibrational modes with negative mode Grüneisen parameters contribute to the NTE of GaFe(CN)6. Our results in the present cubic GaFe(CN)6 suggest that the “guitar-string” effect can be commonly present also in other NTE framework materials.



(5) Dove, M. T.; Fang, H. Negative Thermal Expansion and Associated Anomalous Physical Properties: Review of the Lattice Dynamics Theoretical Foundation. Rep. Prog. Phys. 2016, 79, 066503. (6) Lind, C. Two Decades of Negative Thermal Expansion Research: Where Do We Stand? Materials 2012, 5, 1125. (7) Jiang, X. X.; Molokeev, M. S.; Gong, P.; Yang, Y.; Wang, W.; Wang, S. H.; Wu, S. F.; Wang, Y. X.; Huang, R. J.; Li, L. F.; Wu, Y. C.; Xing, X. R.; Lin, Z. S. Near-Zero Thermal Expansion and High Ultraviolet Transparency in a Borate Crystal of Zn4B6O13. Adv. Mater. 2016, 28, 7936−7940. (8) Evans, J. S. O.; Mary, T. A.; Sleight, A. W. Negative Thermal Expansion in a Large Molybdate and Tungstate Family. J. Solid State Chem. 1997, 133, 580−583. (9) Mary, T. A.; Sleight, A. W. Bulk Thermal Expansion for Tungstate and Molybdates of the Type A2M3O12. J. Mater. Res. 1999, 14, 912−915. (10) Mittal, R.; Chaplot, S. L. Lattice Dynamical Calculation of Isotropic Negative Thermal Expansion in over 0−1050 K. Phys. Rev. B: Condens. Matter Mater. Phys. 1999, 60, 7234. (11) Tallentire, S. E.; Child, F.; Fall, I.; Vella-Zarb, L.; Evans, I. R.; Tucker, M. G.; Keen, D. A.; Wilson, C.; Evans, J. S. Systematic and Controllable Negative, Zero, and Positive Thermal Expansion in Cubic Zr1‑xSnxMo2O8. J. Am. Chem. Soc. 2013, 135, 12849−12856. (12) Greve, B. K.; Martin, K. L.; Lee, P. L.; Chupas, P. J.; Chapman, K. W.; Wilkinson, A. P. Pronounced Negative Thermal Expansion from a Simple Structure: Cubic ScF3. J. Am. Chem. Soc. 2010, 132, 15496−15498. (13) 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. (14) Hu, L.; Chen, J.; Sanson, A.; Wu, H.; Guglieri Rodriguez, C.; Olivi, L.; Ren, Y.; Fan, L. L.; Deng, J. X.; Xing, X. R. New Insights into the Negative Thermal Expansion: Direct Experimental Evidence for the “Guitar-String” Effect in Cubic ScF3. J. Am. Chem. Soc. 2016, 138, 8320−8323. (15) 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. (16) 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. (17) Margadonna, S.; Prassides, K.; Fitch, A. N. Zero Thermal Expansion in a Prussian Blue Analogue. J. Am. Chem. Soc. 2004, 126, 15390−15391. (18) Goodwin, A. L.; Chapman, K. W.; Kepert, C. J. GuestDependent Negative Thermal Expansion in Nanoporous Prussian Blue Analogues MIIPtIV(CN)6⊙x{H2O}(0 ≤ x ≤ 2; M = Zn, Cd). J. Am. Chem. Soc. 2005, 127, 17980−17981. (19) Goodwin, A. L.; Calleja, M.; Conterio, M. J.; Dove, M. T.; Evans, J. S.; 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. (20) Duyker, S. G.; Peterson, V. K.; Kearley, G. J.; RamirezCuesta, 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. 2013, 125, 5374−5378. (21) 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. (22) Dubbeldam, D.; Walton, K. S.; Ellis, D. E.; Snurr, R. Q. Exceptional Negative Thermal Expansion in Isoreticular MetalOrganic Frameworks. Angew. Chem. 2007, 119, 4580−4583.

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.8b01526.



Data analysis procedures and supporting figure (PDF)

AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Jun Chen: 0000-0002-7330-8976 Author Contributions

J.C. conceived the idea and designed the experiments. Q.G. prepared the samples, performed the measurements, and analysis. Q.S. performed density functional theory (DFT) calculations. A.S., R.M., and A.C. performed the X-ray absorption fine structure study. Y.R., S.H.L., and H.Z. assisted in the synchrotron powder diffraction experiments. Q.H. assisted in the neutron powder diffraction experiments. All authors contributed to the discussions. Q.G. and J.C. wrote the manuscript with the help for the revision from all coauthors. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was supported by the National Natural Science Foundation of China (Grant Nos. 21731001, 21590793), the Fundamental Research Funds for the Central Universities, China (FRF-TP-17-001B), the Changjiang Young Scholars Award, and the National Program for Support of Top-notch Young Professionals. Use of the Advanced Photon Source, an Office of Science User Facility operated for the U.S. Department of Energy (DOE), Office of Science, by Argonne National Laboratory, was supported by the U.S. DOE under Contract No. DE-AC02-06CH11357. We acknowledge ELETTRA Synchrotron Light Laboratory for provision of synchrotron radiation (Experiment 20175297), as well as the staff of XAFS Beamline for technical assistance.



REFERENCES

(1) 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. (2) Mohn, P. Materials Science: A Century of Zero Expansion. Nature 1999, 400, 18−19. (3) Attfield, J. P. Condensed-Matter Physics: A Fresh Twist on Shrinking Materials. Nature 2011, 480, 465−466. (4) 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. F

DOI: 10.1021/acs.inorgchem.8b01526 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

(43) Sayers, D. E.; Bunker, B. A. In X-ray Absorption: Principles, Applications, Techniques of EXAFS, SEXAFS and XANES; Koningsberger, D. C., Prins, R., Eds.; Wiley: New York, 1988. (44) Fornasini, P. In Synchrotron Radiation - Basics, Methods and Applications; Mobilio, S., Meneghini, C., Boscherini, F., Eds.; Springer-Verlag: Berlin, 2015; Chapter 6, pp 181−211. (45) Ankudinov, L.; Ravel, B.; Rehr, J. J.; Conradson, S. D. Realspace Multiple-scattering Calculation and Interpretation of X-rayabsorption Near-edge Structure. Phys. Rev. B: Condens. Matter Mater. Phys. 1998, 58, 7565. (46) Newville, M.; Ravel, B.; Rehr, J. J.; Stern, E.; Yacoby, Y.; Haskel, D. Analysis of Multiple-scattering XAFS Data using Theoretical Standards. Phys. B 1995, 208−209, 154. (47) Kresse, G.; Furthmüller, J. Efficient Iterative Schemes for ab Initio Total-energy Calculations using a Plane-wave Basis Set. Phys. Rev. B: Condens. Matter Mater. Phys. 1996, 54, 11169. (48) Kresse, G.; Joubert, D. From Ultrasoft Pseudopotentials to the Projector Augmented-wave Method. Phys. Rev. B: Condens. Matter Mater. Phys. 1999, 59, 1758. (49) Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865. (50) Togo, A.; Oba, F.; Tanaka, I. First-principles Calculations of the Ferroelastic Transition between Rutile-type and CaCl2-type SiO2 at high pressures. Phys. Rev. B: Condens. Matter Mater. Phys. 2008, 78, 134106. (51) Adak, S.; Daemen, L. L.; Nakotte, H. Negative Thermal Expansion in the Prussian Blue Analogues Zn3[Fe(CN)6]2: X-Ray Diffraction and Neutron Vibrational Studies. J. Phys. Conf. Ser. 2010, 251, 012007. (52) Tokoro, H.; Nakagawa, K.; Imoto, K.; Hakoe, F.; Ohkoshi, S. I. Zero Thermal Expansion Fluid and Oriented Film Based on a Bistable Metal-Cyanide Polymer. Chem. Mater. 2012, 24, 1324−1330. (53) Kishore, D.; Gupta, G. P.; Lal, K. C.; Srivastava, T. N. Mössbauer Spectra of In(III) and Ga(III) Hexacyanoferrate(III). Z. Phys. Chem. 1981, 126, 127−128. (54) Haghighi, B.; Khosravi, M.; Barati, A. Fabrication of Gallium Hexacyanoferrate Modified Carbon Ionic Liquid Paste Electrode for Sensitive Determination of Hydrogen Peroxide and Glucose. Mater. Sci. Eng., C 2014, 40, 204−211. (55) Eftekhari, A. Electrochemical Behavior of Gallium Hexacyanoferrate Film Directly Modified Electrode in a Cool Environment. J. Electrochem. Soc. 2004, 151, E297−E301. (56) Katafias, A.; Impert, O.; Kita, P. Hydrogen Peroxide as a Reductant of Hexacyanoferrate(III) in Alkaline Solutions: Kinetic Studies. Transition Met. Chem. 2008, 33, 1041−1046. (57) Fornasini, P.; Grisenti, R. On EXAFS Debye-Waller Factor and Recent Advances. J. Synchrotron Radiat. 2015, 22, 1242. (58) Vaccari, M.; Grisenti, R.; Fornasini, P.; Rocca, F.; Sanson, A. Negative Thermal Expansion in Cucl: An Extended X-Ray Absorption Fine Structure Study. Phys. Rev. B: Condens. Matter Mater. Phys. 2007, 75, 184307.

(23) Han, S. S.; Goddard, W. A. Metal-Organic Frameworks Provide Large Negative Thermal Expansion Behavior. J. Phys. Chem. C 2007, 111, 15185−15191. (24) 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. (25) Zhou, H. L.; Zhang, Y. B.; Zhang, J. P.; Chen, X. M. Supramolecular-Jack-Like Guest in Ultramicroporous Crystal for Exceptional Thermal Expansion Behavior. Nat. Commun. 2015, 6, 6917. (26) Takenaka, K.; Takagi, H. Giant Negative Thermal Expansion in Ge-Doped Anti-Perovskite Manganese Nitrides. Appl. Phys. Lett. 2005, 87, 261902. (27) Chen, J.; Wang, F. F.; Huang, Q. Z.; Hu, L.; Song, X. P.; Deng, J. X.; Yu, Y. B.; Xing, X. R. Effectively Control Negative Thermal Expansion of Single-Phase Ferroelectrics of PbTiO3-(Bi,La)FeO3 over a Giant Range. Sci. Rep. 2013, 3, 2458. (28) Chen, J.; Xing, X. R.; Sun, C.; Hu, P. H.; Yu, R. B.; Wang, X. W.; Li, L. H. Zero Thermal Expansion in PbTiO3-Based Perovskites. J. Am. Chem. Soc. 2008, 130, 1144−1145. (29) Azuma, M.; Chen, W. T.; Seki, H.; Czapski, M.; Oka, K.; Mizumaki, M.; Watanuki, T.; Ishimatsu, N.; Kawamura, N.; Ishiwata, S.; Tucker, M. G.; Shimakawa, Y.; Attfield, J. P.; Olga, S. Colossal Negative Thermal Expansion in BiNiO3 Induced by Intermetallic Charge Transfer. Nat. Commun. 2011, 2, 347. (30) Long, Y. W.; Hayashi, V.; Saito, T.; Azuma, M.; Muranaka, S.; Shimakawa, Y. Temperature-Induced A−B Intersite Charge Transfer in an A-site-ordered LaCu3Fe4O12 Perovskite. Nature 2009, 458, 60. (31) Braden, M.; Andre, G.; Nakatsuji, S.; Maeno, Y. Crystal and Magnetic Structure of Magnetoelastic Coupling and the MetalInsulator Transition. Phys. Rev. B: Condens. Matter Mater. Phys. 1998, 58, 847. (32) Takenaka, K.; Okamoto, Y.; Shinoda, T.; Katayama, N.; Sakai, Y. Colossal Negative Thermal Expansion in Reduced Layered Ruthenate. Nat. Commun. 2017, 8, 14102. (33) Tucker, M. G.; Goodwin, A. L.; Dove, M. T.; Keen, D. A.; Wells, S. A.; Evans, J. S. O. Negative Thermal Expansion in: Mechanisms, Rigid Unit Modes, and Neutron Total Scattering. Phys. Rev. Lett. 2005, 95, 255501. (34) Cao, D.; Bridges, F.; Kowach, G. R.; Ramirez, A. P. Frustrated Soft Modes and Negative Thermal Expansion in ZrW2O8. Phys. Rev. Lett. 2002, 89, 215902. (35) Sanson, A. Toward an Understanding of the Local Origin of Negative Thermal Expansion in ZrW2O8: Limits and Inconsistencies of the Tent and Rigid Unit Mode Models. Chem. Mater. 2014, 26, 3716−3720. (36) Ohkoshi, S.; Tokoro, H.; Matsuda, T.; Takahashi, H.; Irie, H.; Hashimoto, K. Coexistence of Ferroelectricity and Ferromagnetism in a Rubidium Manganese Hexacyanoferrate. Angew. Chem. 2007, 119, 3302−3305. (37) Kaye, S. S.; Long, J. R. Hydrogen Storage in the Dehydrated Prussian Blue Analogues M3[Co(CN)6]2 (M = Mn, Fe, Co, Ni, Cu, Zn). J. Am. Chem. Soc. 2005, 127, 6506−6507. (38) Gao, Q. L.; Chen, J.; Li, Q.; Zhang, J.; Zhai, Z.; Zhang, S. T.; Yu, R. B.; Xing, X. R. Structure and Excellent Visible Light Catalysis of Prussian Blue Analogues BiFe(CN)6·4H2O. Inorg. Chem. Front. 2018, 5, 438−445. (39) Patra, C. R. Prussian Blue Nanoparticles and Their Analogues for Application to Cancer Theranostics. Nanomedicine 2016, 11, 569. (40) Rodríguez-Carvajal, J. Recent Advances in Magnetic Structure Determination by Neutron Powder Diffraction. Phys. B 1993, 192, 55−69. (41) Qiu, X.; Thompson, J. W.; Billinge, S. J. PDFgetX2: a GUIdriven program to obtain the pair distribution function from X-ray powder diffraction data. J. Appl. Crystallogr. 2004, 37, 678. (42) Bridges, F.; Keiber, T.; Juhas, P.; Billinge, S. J. L.; Sutton, L.; Wilde, J.; Kowach, G. R. Local vibrations and negative thermal expansion in ZrW2O8. Phys. Rev. Lett. 2014, 112, 045505. G

DOI: 10.1021/acs.inorgchem.8b01526 Inorg. Chem. XXXX, XXX, XXX−XXX