• No results found

Influence of Metal Defects on the Mechanical Properties of ABX(3) Perovskite-Type Metal-formate Frameworks

N/A
N/A
Protected

Academic year: 2021

Share "Influence of Metal Defects on the Mechanical Properties of ABX(3) Perovskite-Type Metal-formate Frameworks"

Copied!
5
0
0

Loading.... (view fulltext now)

Full text

(1)

In fluence of Metal Defects on the Mechanical Properties of ABX 3

Perovskite-Type Metal-formate Frameworks

Hanna L. B. Boström* and Gregor Kieslich*

Cite This:J. Phys. Chem. C 2021, 125, 1467−1471 Read Online

ACCESS

Metrics & More Article Recommendations

*

sı Supporting Information

ABSTRACT: Defects are emerging as a key tool for fine-tuning the stimuli-responsive behavior of coordination polymers and metal −organic frameworks. Here, we study the rami fications of defects on the mechanical properties of the molecular perovskite [C(NH

2

)

3

]Mn

II

(HCOO)

3

and its defective analogue [C(NH

2

)

3

]Fe

2/3III

1/3

(HCOO)

3

, where □ = vacancy. Defects reduce the bulk modulus by 30% and give rise to a temperature- driven phase transition not observed in the nondefective system. The results highlight the opportunities that come with defect-engineering approaches to alter the mechanical properties and underlying thermodynamics, with important implications for the research on stimuli-responsive materials.

■ INTRODUCTION

Molecular perovskites are dense coordination networks with an ABX

3

perovskite-type structure, where A and/or X is a molecular ion. These compounds include a wealth of compositions such as formates,

1,2

azides,

3,4

thiocyanates,

5,6

dicyanamides,

7,8

dicyano- metallates,

9,10

and, conceptually related, Prussian blue ana- logues, covering phenomena of fundamental and technological relevance.

11,12

Compared to their inorganic counterparts, molecular perovskites show additional degrees of chemical and structural freedom as enabled by the presence of polyatomic species.

13,14

For instance, the use of polyatomic A-site cations promotes the occurrence of temperature- and pressure-driven phase changes as a result of order −disorder transitions of the molecular A-site cation, opening a variety of functionality to be exploited.

15,16

Looking at ways to fine-tune material properties, the incorporation of defects have proved to be a tremendously successful approach in porous coordination polymers.

17

For example, the archetypical defective metal −organic framework UiO-66 supports both missing linkers and missing clusters,

18,19

which has implications for e.g. the adsorption properties.

20,21

However, good mechanical properties are a prerequisite for a material to be practically useful and hence there is a strong need to elucidate the interplay between defects and the nonambient behavior.

22

Several studies have been focused on this issue, using both experimental and computational techniques,

19,23,24

but further investigation is still required. Concerning molecular perovskites, the incorporation of a large concentration of Schottky-type metal defects has been achieved recently,

25

and its impact on the stimuli-responsive properties is still unknown.

Here, we study the bulk modulus (B

0

) and volumetric coe fficient of thermal expansion (α

V

) of the metal-formate framework [C(NH

2

)

3

]Mn

II

(HCOO)

3

and its defective ana- logue [C(NH

2

)

3

]Fe

2/3III

1/3

(HCOO)

3

( □ = vacancy and

C(NH

2

)

3

= guanidinium) using powder X-ray di ffraction (PXRD) under variable temperature and pressure. For simplicity, these will be referred to as GuaMn and GuaFe

2/3

1/3

. We note that these quantities also serve as probes of the Gibbs free energy surface through the relationships

( )

B V p

G T 0

2

= − 2

and

α =V V1

( )

∂ ∂p T2G p

, such that our results provide qualitative insight into the free energy surface of these materials.

METHODS

All chemicals were used as purchased without further puri fication. The samples were synthesized by previously reported methods.

26,27

A solution of [C(NH

2

)

3

]

2

CO

3

(281.5 mg, 1.56 mmol) and HCOOH (182.5 μL, 4.84 mmol) in methanol (5 mL) was added to a methanolic solution of 0.1 M Mn(NO

3

)

2

or 0.1 M FeCl

3

(5 mL). The reaction mixture was stirred overnight at room temperature prior to isolation by filtration in vacuo and washing with methanol.

Variable-temperature PXRD data of GuaFe

2/3

1/3

were collected using a STOE stadi P with Mo K α1 radiation and an Oxford di ffraction Cryosystem. Variable-pressure PXRD data of both GuaMn and GuaFe

2/3

1/3

were collected using a home- built setup at beamline I15 with λ = 0.4246 Å, Diamond Light Source, UK.

28,29

Received: October 30, 2020 Revised: November 25, 2020 Published: January 4, 2021

Article pubs.acs.org/JPCC

License, which permits unrestricted use, distribution and reproduction in any medium, provided the author and source are cited.

Downloaded via UPPSALA UNIV on March 12, 2021 at 15:24:49 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

(2)

Data analysis was carried out by Pawley re finement as implemented in the software TOPAS.

30,31

The peakshape was modelled by a pseudo-Voigt function and the background by a Chebyshev polynomial. The variable-pressure unit cell lattice parameters were fitted using the second-order Birch−

Murnaghan equation of state as implemented in the software EoS fit-GUI.

32−34

■ RESULTS AND DISCUSSION

GuaMn and GuaFe

2/3

1/3

are topologically identical, but di ffer in the metal occupancy and symmetry. GuaMn crystallizes in the space group Pnna,

26

driven by R-point octahedral tilting (a

0

b

b

in Glazer notation

35,36

) and the orientational X-point order of the guanidinium cations [Figure 1a].

25

Aliovalent substitution of Mn

II

in GuaMn with Fe

III

introduces metal vacancies as a charge compensation mechanism, leading to the composition GuaFe

2/3

1/3

[Figure 1b].

27

Despite the large vacancy concentration, structural integrity in GuaFe

2/3

1/3

is retained with only a small lattice strain, on account of the strong hydrogen bonds between the guanidinium and the formate moieties.

37

Given the energy penalty associated with next- neighbor vacancy pairs, the vacancies partially order, which reduces the symmetry to the monoclinic space group P2/n11 ( ≡P12/c1) with α ∼ 89.5°.

27

The nonstandard setting of the monoclinic space group allows for direct comparison of lattice parameters between orthorhombic GuaMn and monoclinic GuaFe

2/3

1/3

. Notably, coordination networks are known to harbor a wealth of defects, yet missing ion defects are still relatively rare.

17,20,38,39

Variable-pressure di ffraction was meas- ured up to 0.4 GPa at I15, Diamond Light Source, using equipment dedicated to accurate control of low pressures.

28

The bulk moduli were calculated using EoSFit and compressibilities by linear fits [ Table 1, Figures S1 −S4 ].

32

The bulk modulus of GuaMn is 20.0(3) GPa, which is in good agreement with the value obtained over a larger pressure range:

B

0

= 21.3 GPa.

41

Compared to other formate-based perovskites, this value is slightly lower than for dimethylammonium metal formates (B

0

= 21.3 GPa for M = Mn

II

and 26.3 GPa for M = Co

II

),

16,42,43

but similar to that of [NH

2

NH

3

]Zn(HCOO)

3

with B

0

= 19 GPa.

44

The defective GuaFe

2/3

1/3

features a lower

bulk modulus of 14.3(2) GPa, which is comparable to the behavior of the heterometallic [C

2

H

5

NH

3

]K

0.5

Al

0.5

(HCOO)

3

.

45

The di fference in bulk moduli of GuaMn and defective GuaFe

2/3

1/3

highlights how defects increase the mechanical compliance of the framework. Importantly, the impact of defects even overcompensates in fluence of ion size, which is known to make molecular perovskites more robust.

40

Both systems show an anisotropic pressure response with K

a

>

K

b

> K

c

, where K

i

is the compressibility of axis i [Table 1, Figure 1d]. To compare the compressibilities of the two systems, the mechanical building unit approach was employed.

46

The lattice parameters for both systems can be recast as

Ä ÇÅÅÅÅÅ ÅÅÅ

É ÖÑÑÑÑÑ ÑÑÑ

r b

a c

1 3 2

2 2

̅ = + +

(1)

and

i kjjj yc{zzz 2 tan 1 a

θ =

(2)

where r ̅ is the average strut length, that is, the metal−metal distance, and θ is the hinging angle, that is, metal−metal−metal angle within the ac plane [Figure 1c].

40

Here, θ > 90°. This approach represents a physically meaningful parametrization of

Figure 1.Structures (a) GuaFe2/31/3and (b) GuaMn26with the unit cell indicated in black. For GuaFe2/31/3, transparent octahedra are 1/3 occupied. (c) Metal framework with the hinging angle (θ) and one of the strut lengths (r) indicated. The pressure-induced percentage changes in (d) cell dimensions, (e) average strut length r̅, and (f) hinging angle θ for GuaMn and GuaFe2/31/3. Errors are smaller than the size of the data markers.

Table 1. Volumetric and Linear Bulk Moduli and Expansivities for GuaMn and GuaFe

2/31/3a

Mn Fe

Ka/TPa−1 30.1(4) 36.7(2)

Kb/TPa−1 17.8(3) 25.8(2)

Kc/TPa−1 0.07(14) 6.9(3)

B0/GPa 20.0(3) 14.3(2)

αa/MK−1 39(2) 35(3)

αb/MK−1 30(1) 27(3)

αc/MK−1 −5.4(5) −8(2)

αV/MK−1 63(2) 58(6)

aThermal expansion was fitted in the range 300−230 K and compressibilities up to 400 MPa. Variable-temperature data for GuaMn are taken from ref40.

(3)

the structural changes upon compression, highlighting the behavior of the pseudocubic 3D network. As expected, the strut length r ̅ decreases with pressure, which is accompanied by an increase in the hinging of the framework [Figure 1e,f].

The behaviors of r ̅ and θ upon compression rationalize the anisotropic compression. As given in eqs 1 and 2, the cell dimensions are linear functions of r ̅, but with varying dependence on θ. While c is positively correlated with θ, a is negatively correlated and b is independent. Thus, a is the most compliant direction, as both the decrease of r ̅ and the increase of θ upon compression reduce its length. The converse scenario appears for c: the variations of r ̅ and θ operate in tension, which gives a low value of K

c

. Chemically, the low expansion can be explained by recognizing that the c axis runs along the planes of the incompressible guanidinium cations.

41

In GuaMn, the opposing e ffects of r̅ and θ on K

c

are equal in magnitude and cancel, leading to zero uniaxial compressibility. Further increase of the hinging mechanism may lead to negative linear compressibility. The defects in GuaFe

2/3

1/3

soften r ̅, as a result of the larger void space, whereas the pressure-induced increase of θ is unchanged. This does not substantially impact the compressibilities of a and c, but K

c

increases to a nonzero value.

Furthermore, the thermal behavior of GuaFe

2/3

1/3

was investigated in the range 100 −300 K by variable-temperature PXRD. The system remains monoclinic down to 220 K, where additional re flections emerge in the diffraction patterns [ Figures S5 and S6]. Because of the low symmetry and lack of high- resolution synchrotron data, the low-temperature phase was not solved and Pawley re finements

30

were only carried out for the ambient phase. The phase transition in defective GuaFe

2/3

1/3

contrasts with the variable-temperature behavior of GuaMn reported by Collings et al., where the system remained in its ambient phase down to 100 K.

40

Hence, the metal defects alter the phase behavior, which may possibly be driven by the enhancement of flexibility by virtue of the larger free space. It is noteworthy that no phase transition was observed in the variable-pressure XRD data, although it is possible that the signal was not su fficiently strong. Further studies into the low- temperature phase would be of interest.

The volumetric and linear expansivities ( α

i

for axis i) were calculated in the range 300 −230 K by PASCal [ Table 1, Figure 2].

47

GuaMn and GuaFe

2/3

1/3

show equal (within error) volumetric coe fficients of thermal expansion and both systems are highly anisotropic. Mirroring the pressure behavior, the linear expansivities decrease in the order |α

a

| > |α

b

| > |α

c

| ( Figure 2 and Table 1). Notably, both compounds exhibit negative thermal expansion (NTE) along the c axis. While rare in purely

inorganic compounds, NTE is relatively common amongst formate-based perovskites

40,48

and coordination networks in general.

49,50

It appears that the defects play a less important role in the thermal behavior compared to the pressure response, but it should be noted that the investigated temperature range is relatively small.

Defects clearly modify the mechanical properties of GuaFe

2/3

1/3

relative to GuaMn; removing 1/3 of the transition metals reduces the bulk modulus by ∼30%. With the emerging understanding of defect chemistry in MOFs and coordination polymers, there is also growing interest in the role of defects on mechanical properties. It can be misleading to compare results between di fferent systems, as the results are likely dependent on the precise topologies and type of defects.

Additionally, solvent content strongly impacts the mechanical response, which may complicate comparisons.

51

However, we note that defects reduce the bulk modulus of UiO-66 up to a critical concentration and certain Prussian blue analogues, in line with the results presented here.

24,52

This likely results from the increased void space, which correlates with greater ease of compression.

53

Comparisons of similar studies may enable the development of general guidelines regarding the in fluence of defects on the mechanical properties in dense and porous coordination polymers.

■ CONCLUSIONS

To conclude, we report the stimuli-responsive behavior of defect-free [C(NH

2

)

3

]Mn

I I

(HCOO)

3

and defective [C(NH

2

)

3

]Fe

2/3III

1/3

(HCOO)

3

, and establish how missing- ion vacancies control the structural response to pressure and temperature variation in molecular perovskites for the first time.

Upon compression, defects selectively soften the average strut length r ̅, without having an effect on the hinging angle θ. The defective GuaFe

2/3

1/3

undergoes a thermal phase transition not observed in GuaMn, yet both systems exhibit similar thermal expansion behavior. Placing our results in the context of e fforts to manipulate the free energy landscape in coordination polymers, we stress that using defects as a tool to access (stimuli-responsive) properties is an increasingly appealing strategy. For instance, barocaloric behavior has recently been observed in certain molecular perovskites

54

and defects o ffer intriguing opportunities for controlling the underlying thermo- dynamics to optimize the barocaloric performance. In particular, defects flatten the underlying free energy surface, which is expected to impact phase transition temperatures. This suggests that defects can be used to enhance the barocaloric properties, for example, by optimizing the temperature of the barocaloric phase transition. More broadly, it will be exciting to see how defects will be exploited to learn about the fundamental interactions in coordination networks. The bulk modulus appears to be an important probe for the impact of defects on the underlying chemical interactions as re flected in the free energy landscape.

■ ASSOCIATED CONTENT

*

Supporting Information

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.jpcc.0c09796.

Experimental details, XRD patterns, evolution of lattice parameters as a function of temperature and pressure, and f −F plots ( PDF)

Figure 2.Percentage change in the unit cell volume as a function of temperature for GuaMn (filled squares) and GuaFe2/31/3 (empty diamonds).

(4)

■ AUTHOR INFORMATION

Corresponding Authors

Hanna L. B. Bostro ̈m − Max Planck Institute for Solid State Research, D-70569 Stuttgart, Germany; Department of Inorganic Chemistry, Ångström Laboratory, Uppsala University, SE-751 21 Uppsala, Sweden; orcid.org/0000- 0002-8804-298X; Email: h.bostroem@fkf.mpg.de Gregor Kieslich − Department of Chemistry, Technical

University of Munich, D-85748 Garching, Germany;

orcid.org/0000-0003-2038-186X;

Email: gregor.kieslich@tum.de

Complete contact information is available at:

https://pubs.acs.org/10.1021/acs.jpcc.0c09796

Notes

The authors declare no competing financial interest.

■ ACKNOWLEDGMENTS

The authors thank Diamond Light Source for beamtime CY22477-2 at I15 and Stefan Burger, Dominik Daisenberger, Karina Hemmer, and Pia Vervoorts for assistance with data collection.

(1) Sletten, E.; Jensen, L. H. The Crystal Structure of Dimethy-

REFERENCES

lammonium Copper(II) Formate, NH2(CH3)2[Cu(OOCH)3]. Acta Crystallogr., Sect. B: Struct. Crystallogr. Cryst. Chem. 1973, 29, 1752−

1756.

(2) Wang, X.-Y.; Gan, L.; Zhang, S.-W.; Gao, S. Perovskite-like Metal Formates with Weak Ferromagnetism and as Precursors to Amorphous Materials. Inorg. Chem. 2004, 43, 4615−4625.

(3) Mautner, F. A.; Krischner, H.; Kratky, C. Tetramethylammonium- Calcium-Triazid. Darstellung und Kristallstruktur. Monatsh. Chem.

1988, 119, 1245−1249.

(4) Zhao, X.-H.; Huang, X.-C.; Zhang, S.-L.; Shao, D.; Wei, H.-Y.;

Wang, X.-Y. Cation-Dependent Magnetic Ordering and Room- Temperature Bistability in Azido-Bridged Perovskite-Type Com- pounds. J. Am. Chem. Soc. 2013, 135, 16006−16009.

(5) Thiele, G.; Messer, D. S-Thiocyanato- und N-Isothiocyanato- Bindungsisomerie in den Kristallstrukturen von RbCd(SCN)3 und CsCd(SCN)3. Z. Anorg. Allg. Chem. 1980, 464, 255−267.

(6) Cliffe, M. J.; Keyzer, E. N.; Dunstan, M. T.; Ahmad, S.; De Volder, M. F. L.; Deschler, F.; Morris, A. J.; Grey, C. P. Strongly coloured thiocyanate frameworks with perovskite-analogue structures. Chem. Sci.

2019, 10, 793−801.

(7) Tong, M.-L.; Ru, J.; Wu, Y.-M.; Chen, X.-M.; Chang, H.-C.;

Mochizuki, K.; Kitagawa, S. Cation-templated construction of three- dimensional α-Po cubic-type [M(dca)3] networks. Syntheses, structures and magnetic properties of A[M(dca)3) (dca = dicyanamide;

for A = benzyltributylammonium, M = Mn2+, Co2+; for A = benzyltriethylammonium, M = Mn2+, Fe2+. New J. Chem. 2003, 27, 779−782.

(8) Schlueter, J. A.; Manson, J. L.; Hyzer, K. A.; Geiser, U. Spin Canting in the 3D Anionic Dicyanamide Structure (SPh3)Mn(dca)3 (Ph = phenyl, dca = dicyanamide). Inorg. Chem. 2004, 43, 4100−4102.

(9) Hill, J. A.; Thompson, A. L.; Goodwin, A. L. Dicyanometallates as model extended frameworks. J. Am. Chem. Soc. 2016, 138, 5886−5896.

(10) Lefebvre, J.; Chartrand, D.; Leznoff, D. B. Synthesis, structure and magnetic properties of 2-D and 3-D [cation] {M[Au(CN)2]3} (M

= Ni, Co) coordination polymers. Polyhedron 2007, 26, 2189−2199.

(11) Kieslich, G.; Goodwin, A. L. The same and not the same:

molecular perovskites and their solid-state analogues. Mater. Horiz.

2017, 4, 362−366.

(12) Li, W.; Wang, Z.; Deschler, F.; Gao, S.; Friend, R. H.; Cheetham, A. K. Chemically diverse and multifunctional hybrid organic-inorganic perovskites. Nat. Rev. Mater. 2017, 2, 16099.

(13) Boström, H. L. B.; Senn, M. S.; Goodwin, A. L. Recipes for improper ferroelectricity in molecular perovskites. Nat. Commun. 2018, 9, 2380.

(14) Xu, W.-J.; Du, Z.-Y.; Zhang, W.-X.; Chen, X.-M. Structural phase transitions in perovskite compounds based on diatomic or multiatomic bridges. CrystEngComm 2016, 18, 7915−7928.

(15) Jain, P.; Dalal, N. S.; Toby, B. H.; Kroto, H. W.; Cheetham, A. K.

Order-Disorder Antiferroelectric Phase Transition in a Hybrid Inorganic-Organic Framework with the Perovskite Architecture. J.

Am. Chem. Soc. 2008, 130, 10450−10451.

(16) Collings, I. E.; Bykov, M.; Bykova, E.; Hanfland, M.; van Smaalen, S.; Dubrovinsky, L.; Dubrovinskaia, N. Disorder−order transitions in the perovskite metal−organic frameworks [(CH3)2NH2] [M- (HCOO)3] at high pressure. CrystEngComm 2018, 20, 3512−3521.

(17) Dissegna, S.; Epp, K.; Heinz, W. R.; Kieslich, G.; Fischer, R. A.

Defective Metal−Organic Frameworks. Adv. Mater. 2018, 30, 1704501.

(18) Cavka, J. H.; Jakobsen, S.; Olsbye, U.; Guillou, N.; Lamberti, C.;

Bordiga, S.; Lillerud, K. P. A New Zirconium Inorganic Building Brick Forming Metal Organic Frameworks with Exceptional Stability. J. Am.

Chem. Soc. 2008, 130, 13850−13851.

(19) Cliffe, M. J.; Hill, J. A.; Murray, C. A.; Coudert, F.-X.; Goodwin, A. L. Defect-dependent colossal negative thermal expansion in UiO- 66(Hf) metal−organic framework. Phys. Chem. Chem. Phys. 2015, 17, 11586−11592.

(20) Shearer, G. C.; Chavan, S.; Bordiga, S.; Svelle, S.; Olsbye, U.;

Lillerud, K. P. Defect Engineering: Tuning the Porosity and Composition of the Metal-Organic Framework UiO-66 via Modulated Synthesis. Chem. Mater. 2016, 28, 3749−3761.

(21) Liang, W.; Coghlan, C. J.; Ragon, F.; Rubio-Martinez, M.;

D’Alessandro, D. M.; Babarao, R. Defect engineering of UiO-66 for CO2and H2O uptake A combined experimental and simulation study. Dalton Trans. 2016, 45, 4496−4500.

(22) 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.

(23) Rogge, S. M. J.; Wieme, J.; Vanduyfhuys, L.; Vandenbrande, S.;

Maurin, G.; Verstraelen, T.; Waroquier, M.; Van Speybroeck, V.

Thermodynamic Insight in the High-Pressure Behavior of UiO-66:

Effect of Linker Defects and Linker Expansion. Chem. Mater. 2016, 28, 5721−5732.

(24) Dissegna, S.; Vervoorts, P.; Hobday, C. L.; Düren, T.;

Daisenberger, D.; Smith, A. J.; Fischer, R. A.; Kieslich, G. Tuning the Mechanical Response of Metal-Organic Frameworks by Defect Engineering. J. Am. Chem. Soc. 2018, 140, 11581−11584.

(25) Boström, H. L. B. Tilts and shifts in molecular perovskites.

CrystEngComm 2020, 22, 961−968.

(26) Hu, K.-L.; Kurmoo, M.; Wang, Z.; Gao, S. Metal-Organic Perovskites: Synthesis, Structures, and Magnetic Properties of [C- (NH2)3] [MII(HCOO)3] (M = Mn, Fe, Co, Ni, Cu, and Zn; C(NH2)3

= Guanidinium). Chem. Eur. J. 2009, 15, 12050−12064.

(27) Boström, H. L. B.; Bruckmoser, J.; Goodwin, A. L. Ordered B- Site Vacancies in an ABX3Formate Perovskite. J. Am. Chem. Soc. 2019, 141, 17978−17982.

(28) Brooks, N. J.; Gauthe, B. L. L. E.; Terrill, N. J.; Rogers, S. E.;

Templer, R. H.; Ces, O.; Seddon, J. M. Automated high pressure cell for pressure jump X-ray diffraction. Rev. Sci. Instrum. 2010, 81, 064103.

(29) Vervoorts, P.; Keupp, J.; Schneemann, A.; Hobday, C. L.;

Daisenberger, D.; Fischer, R. A.; Schmid, R.; Kieslich, G. Configura- tional Entropy Driven High-Pressure Behaviour of a Flexible Metal−

Organic Framework. Angew. Chem., Int. Ed. 2020, DOI: 10.1002/

anie.202011004.

(30) Pawley, G. S. Unit-cell refinement from powder diffraction scans.

J. Appl. Crystallogr. 1981, 14, 357−361.

(31) Coelho, A. A. TOPAS and TOPAS-Academic: an optimization program integrating computer algebra and crystallographic objects written in C++. J. Appl. Crystallogr. 2018, 51, 210−218.

(32) Angel, R. J.; Alvaro, M.; Gonzalez-Platas, J. EosFit7c and a Fortran module (library) for equation of state calculations. Z.

Kristallogr.-Cryst. Mater. 2014, 229, 405−419.

(5)

(33) Birch, F. Finite elastic strain of cubic crystals. Phys. Rev. 1947, 71, 809−824.

(34) Murnaghan, F. D. The compressibility of media under extreme pressures. Proc. Natl. Acad. Sci. U.S.A. 1944, 30, 244−247.

(35) Glazer, A. M. The Classification of Tilted Octahedra in Perovskites. Acta Crystallogr., Sect. B: Struct. Crystallogr. Cryst. Chem.

1972, 28, 3384−3392.

(36) Howard, C. J.; Stokes, H. T. Group-Theoretical Analysis of Octahedral Tilting in Perovskites. Acta Crystallogr., Sect. B: Struct. Sci.

1998, 54, 782−789.

(37) Svane, K. L.; Forse, A. C.; Grey, C. P.; Kieslich, G.; Cheetham, A.

K.; Walsh, A.; Butler, K. T. How Strong is the Hydrogen Bond in Hybrid Perovskites? J. Phys. Chem. Lett. 2017, 8, 6154−6159.

(38) Sholl, D. S.; Lively, R. P. Defects in Metal−Organic Frameworks:

Challenge or Opportunity? J. Phys. Chem. Lett., 2015, 6, 3437−3444.

(39) Bennett, T. D.; Cheetham, A. K.; Fuchs, A. H.; Coudert, F.-X.

Interplay between defects, disorder and flexibility in metal-organic frameworks. Nat. Chem. 2017, 9, 11−16.

(40) Collings, I. E.; Hill, J. A.; Cairns, A. B.; Cooper, R. I.; Thompson, A. L.; Parker, J. E.; Tang, C. C.; Goodwin, A. L. Compositional Dependence of Anomalous Thermal Expansion in Perovskite-Like ABX3Formates. Dalton Trans. 2016, 45, 4169−4178.

(41) Yang, Z.; Cai, G.; Bull, C. L.; Tucker, M. G.; Dove, M. T.;

Friedrich, A.; Phillips, A. E. Hydrogen-bond-mediated structural variation of metal guanidinium formate hybrid perovskites under pressure. Philos. Trans. R. Soc., A 2019, 377, 20180227.

(42) Sobczak, S.; Chitnis, A.; Andrzejewski, M.; Mączka, M.; Gohil, S.;

Garg, N.; Katrusiak, A. Framework and Coordination Strain in Two Isostructural Hybrid Metal-Organic Perovskites. CrystEngComm 2018, 20, 5348−5355.

(43) Collings, I. E.; Saines, P. J.; Mikolasek, M.; Boffa Ballaran, T.;

Hanfland, M. Static disorder in a perovskite mixed-valence metal− organic framework. CrystEngComm 2020, 22, 2859−2865.

(44) Kieslich, G.; Forse, A. C.; Sun, S.; Butler, K. T.; Kumagai, S.; Wu, Y.; Warren, M. R.; Walsh, A.; Grey, C. P.; Cheetham, A. K. Role of Amine−Cavity Interactions in Determining the Structure and Mechanical Properties of the Ferroelectric Hybrid Perovskite [NH3NH2] Zn(HCOO)3. Chem. Mater. 2016, 28, 312−317.

(45) Ptak, M.; Svane, K. L.; Collings, I. E.; Paraguassu, W. Effect of Alkali and Trivalent Metal Ions on the High-Pressure Phase Transition of [C2H5NH3] M0.5IM0.5III(HCOO)3(MI= Na, K and MIII= Cr, Al) Heterometallic Perovskites. J. Phys. Chem. C 2020, 124, 6337−6348.

(46) Ogborn, J. M.; Collings, I. E.; Moggach, S. A.; Thompson, A. L.;

Goodwin, A. L. Supramolecular mechanics in a metal−organic framework. Chem. Sci. 2012, 3, 3011−3017.

(47) 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.

(48) Li, W.; Thirumurugan, A.; Barton, P. T.; Lin, Z.; Henke, S.;

Yeung, H. H.-M.; Wharmby, M. T.; Bithell, E. G.; Howard, C. J.;

Cheetham, A. K. Mechanical tunability via hydrogen bonding in metal- organic frameworks with the perovskite architecture. J. Am. Chem. Soc.

2014, 136, 7801−7804.

(49) Lock, N.; Wu, Y.; Christensen, M.; Cameron, L. J.; Peterson, V.

K.; Bridgeman, A. J.; Kepert, C. J.; Iversen, B. B. Elucidating negative thermal expansion in MOF-5. J. Phys. Chem. C 2010, 114, 16181− 16186.

(50) 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(1,3,5- benzenetricarboxylate)2. Angew. Chem., Int. Ed. 2008, 47, 8929−8932.

(51) Graham, A. J.; Tan, J.-C.; Allan, D. R.; Moggach, S. A. The effect of pressure on Cu-btc: Framework compression vs. guest inclusion.

Chem. Commun. 2012, 48, 1535−1537.

(52) Boström, H. L. B.; Collings, I. E.; Cairns, A. B.; Romao, C. P.;

Goodwin, A. L. High-pressure behavior of Prussian blue analogues:

interplay of hydration, Jahn-Teller distortions and vacancies. Dalton Trans. 2019, 48, 1647−1655.

(53) Redfern, L. R.; Robison, L.; Wasson, M. C.; Goswami, S.; Lyu, J.;

Islamoglu, T.; Chapman, K. W.; Farha, O. K. Porosity Dependence of Compression and Lattice Rigidity in Metal-Organic Framework Series.

J. Am. Chem. Soc. 2019, 141, 4365−4371.

(54) Bermúdez-García, J. M.; Sánchez-Andújar, M.; Señarís- Rodríguez, M. A. A New Playground for Organic−Inorganic Hybrids:

Barocaloric Materials for Pressure-Induced Solid-State Cooling. J. Phys.

Chem. Lett. 2017, 8, 4419−4423.

References

Related documents

46 Konkreta exempel skulle kunna vara främjandeinsatser för affärsänglar/affärsängelnätverk, skapa arenor där aktörer från utbuds- och efterfrågesidan kan mötas eller

Generally, a transition from primary raw materials to recycled materials, along with a change to renewable energy, are the most important actions to reduce greenhouse gas emissions

The aim of the two-week data modeling was to compare the predictive performance of input-output models and no-input models, as well as tuned one-step predictors and tuned

Vi ville försäkra att kollisionstest mot en doodad inte skulle bli allt för arbetssamt för spelmotorn då test mot många doodads skulle kunna behöva göras även efter bortsortering

Over the entire temperature range, calorimetric control recorded neither heat flows caused by phase trans- formations of the first kind (to which the polymorphic transformation

The quasi-experiment is constructed as follows: (1) the stimulus is a business case introducing the respondent as the managing director of a medium-sized industrial firm; (2)

Utifrån det resultat jag kan se från lärarinnorna och främst då folkskolelärarinnorna blir det tydligt att kvinnorna inte var nöjda med sin situation, utan aktivt jobbade för

Growth and Mechanical Properties of Transition Metal Nitrides