Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Compressing arsenic⋯halogen secondary bonds: a high-pressure structural study of arsenic(III) oxide intercalates with ammonium halides

Piotr A. Guńka *a, Sofija Milos b, Martin Ende c, Frederico Alabarse d, Ronald Miletich b and Kamil F. Dziubek b
aFaculty of Chemistry, Warsaw University of Technology, ul. Noakowskiego 3, 00-664 Warszawa, Poland. E-mail: piotr.gunka@pw.edu.pl
bInstitut für Mineralogie und Kristallographie, Universität Wien, Josef-Holaubek-Platz 2, A-1090 Wien, Austria
cInstitute for Geosciences and Geography, Martin-Luther-University Halle-Wittenberg, 06120 Halle (Saale), Germany
dXPRESS Beamline, Elettra Sincrotrone Trieste S.C.p.A, AREA Science Park, I-34149 Basovizza, Trieste, Italy

Received 9th June 2025 , Accepted 28th July 2025

First published on 29th July 2025


Abstract

Crystal structures of arsenic(III) oxide intercalation compounds with ammonium chloride (NH4Cl·As2O3·1/2H2O), ammonium bromide (NH4Br·2As2O3), and ammonium iodide (NH4I·2As2O3) have been determined under high pressure up to 12, 15 and 11 GPa, respectively. No phase transitions have been observed for the investigated compounds. The compression of arsenic⋯halogen secondary bonds, expressed as penetration indices of the bonds, has been shown to be a linear function of unit-cell volume ratio V/V0, where V0 is the unit-cell volume at ambient pressure, with similar slopes as the compression of arsenic⋯oxygen secondary bonds. The behavior of the arsenic coordination number, expressed as a first-order valence entropy coordination number, at high pressures and the stereoactivity of arsenic lone electron pairs in the studied intercalates are the same as in arsenic(III) oxide polymorphs – the former decreases linearly with V/V0, while the latter remains unchanged. The high-pressure study lends further support to the fact that the nature of arsenic⋯halogen and arsenic⋯oxygen secondary bonds is the same.


Introduction

Intercalation compounds result from reversible inclusion, without covalent bonding, of one kind of molecule or ion in a solid matrix of another compound with a layered structure.1 They are widely used, for instance, as electrode materials in batteries or catalysts. Some chalcogenide intercalates exhibit superconducting properties.2,3 Arsenic(III) oxide intercalates are layered compounds, in which electroneutral layers of As2O3 separate sheets of cations and anions. The introduction of ions into the structure changes the conformation of individual As2O3 layers compared to arsenic(III) oxide polymorphs, but the connectivity of atoms within layers remains the same, and there are no covalent bonds between the ions and As2O3 layers. It has recently been reported that intercalates CsBr·2As2O3 and CsBr·As2O3 exhibit very high birefringence. The latter intercalate crystallizes in the polar space group P63mc and has a very strong second harmonic generation response that is 20.5 times stronger than that of KH2PO4.4

Most of the known As2O3 intercalates, with the exception of the intercalate with NaBr, are hexagonal and comprise non-corrugated polar As2O3 layers that exhibit symmetry of layer group P6mm and separate alternating flat layers of cations and anions (structure type P, see Fig. 1).5,6 Anions interact with arsenic(III) lone electron pairs (LEPs) forming the so-called secondary bonds, which are located trans with respect to the primary As–O bonds.5 Cations template the As2O3 layers and are coordinated by oxygen atoms from the layers.7 In the case of KCl and RbCl, hexagonal hydrated intercalates, YKCl and image file: d5ce00593k-t1.tif, have been obtained in addition to anhydrous ones.8,9 Meanwhile, for NH4Cl, only a hydrated intercalate could be obtained so far.10,11 In this compound, there is an additional layer containing disordered water molecules and ammonium cations. It is sandwiched between two layers of chloride anions (see Fig. 1).11


image file: d5ce00593k-f1.tif
Fig. 1 Crystal structures of intercalates NH4Cl·As2O3·1/2H2O (a and c) and NH4Br·2As2O3 (b and d). Views along the crystallographic [001] (a and b) and [120] (c and d) directions. As, O, N, Cl, and Br atoms are depicted as teal, red, blue, light green and brown spheres, respectively. Violet spheres denote N and O atoms of disordered ammonium cations and water molecules. H atoms are omitted for clarity.

The aim of this work is to compare the compressibility and structural changes in the hydrated intercalate NH4Cl·As2O3·1/2H2O (YNH4Cl) and in anhydrous ones NH4Br·2As2O3 (PNH4Br) and NH4I·2As2O3 (PNH4I). Another goal is to compare the pressure evolution of the As⋯halogen secondary bonds with the pressure dependence of As⋯O secondary bonds present in the As2O3 polymorphs.

Experimental and methodological details

High-pressure single-crystal X-ray diffraction

Single crystals of intercalates YNH4Cl, PNH4Br and PNH4I were obtained as described previously.11 ETH-type diamond anvil cells (DACs) were used with pre-indented stainless steel gaskets with a thickness of ca. 120 μm. Diamond culets of 450 μm were used, and 200 μm holes were electrodrilled in the gaskets. Good-quality single crystals of the intercalates with typical dimensions of 80 μm × 60 μm × 15 μm were placed in the gasket holes together with ruby chips used for pressure determination.12 Subsequently, DACs were gas loaded with argon used as a pressure transmitting medium (PTM). Diffraction data were collected at the Xpress beamline of the Elettra Synchrotron Radiation Facility using 0.4957 Å incident wavelength, 80 μm beam diameter and a PILATUS3 S 6M detector (from DECTRIS).13 Raw frames were imported into CrysAlisPRO, which was used for data reduction.14 Crystal structures were refined using SHELXL 2019/3 invoked from Olex2 ver. 1.5 program.15,16 Crystal structure models were analyzed using PLATON.17 3rd order Birch–Murnaghan (BM) equation of state (EoS) was fitted to the experimental p(V) points using the EosFit software.18,19 All unit-cell parameters are given in Tables S1–S3.

Valence entropy coordination number, bond-valence vector, and penetration index calculations

Bond valences (BVs) si were calculated according to the following equation: si = exp((R0di)/B), where di, R0 and B stand for the bond length of bond i, bond length of a bond with a valence of 1 and bond softness parameter, respectively.20 Parameters determined by Gagné and Hawthorne were used for As–O bonds, by Brese and O'Keeffe for As⋯Cl interactions and by Brown for As⋯Br and As⋯I secondary bonds.21–23 The dependence of R0 on pressure was described using the relationship proposed by Brown.24 The magnitude of bond-valence vectors vi (BVVs) was calculated using the relationship derived by Zachara: ‖vi‖ = si(1 − si/Q) where Q is the electric charge of the coordination center atomic core.25 Zachara's BVV definition assumes that it is an integral of the electric field across a surface corresponding to a particular bond irrespective of the direction of that surface. Consequently, BVVs are aligned along the bonds they correspond to; the sum of BVVs around an unstrained coordination center is a zero vector. When a coordination center is electronically strained due to the presence of a stereoactive LEP, the resultant BVV can be brought back to zero by treating the LEP as a pseudoligand with sLEP ≤ 2. The equality holds for a fully stereoactive LEP and values smaller than two indicate partially stereoactive LEPs. When calculating BVVs, arsenic was treated as having a +5 oxidation state, so that the LEP could be treated as a pseudoligand. As no BV parameters for As(V)⋯Br and As(V)⋯I interactions are available, parameters for As(III) were used for these. This is justified by the fact that the difference between BV parameters for As(III)⋯Cl and As(V)⋯Cl interactions is small i.e. R0 values differ by 0.02 Å and B is the same.

The strength of secondary interactions was also gauged using the penetration index defined as: pAB = 100·(vA + vBdAB)/(vA + vBrArB), where vi and ri stand for van der Waals and covalent radii of element i, while pAB represents the penetration index of an interaction with distance dAB between atoms A and B.26 Covalent and van der Waals radii provided by Echeverria and Alvarez were utilized.26

The first-order valence entropy coordination number (1VECN) was applied to quantitatively measure the coordination number of arsenic atoms forming both primary and secondary bonds. It was calculated according to the following formula: image file: d5ce00593k-t2.tif where N is the number of bonds considered and image file: d5ce00593k-t3.tif. More details on the calculation of the 1VECN are given elsewhere.27

Density functional theory (DFT) computations

Equation of state (EoS) modelling by DFT computations was only carried out for the intercalates PNH4Br and PNH4I. The intercalate YNH4Cl was not modelled computationally because there is additional positional disorder in that structure within the layer at z = 1/2. Computations were carried out in a plane-wave basis set using Vienna ab initio simulation package VASP 6.4.2.28 An energy cut-off of 1200 eV was applied with automatic generation of the Monkhorst–Pack net for sampling the first Brillouin zone, and the parameter value defining spacing between the points was set to 40.29 PAW PBE potentials were applied.30 Grimme's D3 correction for dispersion was used together with the Becke–Johnson damping function.31,32 Geometry optimizations were stopped when the energy and maximum force per atom between steps were smaller than 10−7 eV and 0.005 eV Å−1, respectively. Two structural optimizations with only the unit-cell volume frozen followed by the final optimization with only atomic positions allowed to vary were carried out. Ammonium cations located at 0, 0, 0 are disordered and, for the purpose of modelling, we only considered cations with one of N–H bonds aligned along the crystallographic [001] direction.

Results and discussion

Crystal structure evolution with pressure

The pressure dependence of unit-cell volumes of the studied intercalates is plotted in Fig. 2 and the parameters of the fitted 3rd order BM EoS are listed in Table 1. YNH4Cl, PNH4Br and PNH4I undergo smooth compression under high pressure at least up to 12, 15 and 11 GPa, respectively. Apparently, all the investigated crystal structures of intercalates are so robust that argon non-hydrostaticity is only manifested in the increased mosaicity of the studied single crystals (see Fig. S1). Also, the V(p) dependence for intercalates PNH4Br and PNH4I is in very good agreement with DFT predictions computed using the PBE-D3(BJ) model.
image file: d5ce00593k-f2.tif
Fig. 2 Unit-cell volume of the studied intercalates plotted as a function of pressure. Open and closed circles correspond to pressure increase and decrease, respectively. Solid lines correspond to 3rd-order Birch–Murnaghan equations of states fitted to experimental data points, whereas black points correspond to the PBE-D3(BJ) DFT computations described thoroughly in the methodological details section.
Table 1 Parameters of the 3rd-order BM EoS fitted to the experimental and DFT data for the studied intercalates
Y NH 4 Cl P NH 4 Br P NH 4 I
Exp Exp DFT Exp DFT
V 03 304.5(5) 221.3(7) 218.3(6) 226.5(7) 222.5(7)
B 0/GPa 21.3(11) 25(5) 35.2(14) 25(3) 34.1(15)
image file: d5ce00593k-t4.tif 6.8(5) 9(3) 5.8(2) 8(1) 5.82(19)


All the studied intercalates exhibit virtually the same linear compressibility along the <100> crystallographic direction (see Table S4 for ambient-pressure values, linear moduli, and their pressure derivatives of lattice parameters of the studied intercalates and Fig. S2). Intercalates PNH4Br and PNH4I also exhibit very similar linear compressibility along the [001] direction which results in very similar bulk moduli for the two compounds.

Pressure dependence of secondary As⋯halogen bonds

BVs as well as penetration indices of As⋯X secondary bonds are plotted as a function of pressure in Fig. 3. We have shown previously that the strength of As⋯X secondary bonds increases with the size and polarizability of the halide anion.5 Here, the trend is clear for the intercalates PNH4Br and PNH4I, while the intercalate YNH4Cl seems to depart from it at first sight. However, we propose that high values of the As⋯Cl penetration index and BVs result from the different structures of this intercalate. The sAs⋯Cl BV values are 0.025, 0.026 and 0.024 v.u. for PKCl, PRbCl and PCsCl, respectively and 0.041, 0.041 and 0.042 for YNH4Cl, YKCl and image file: d5ce00593k-t5.tif, respectively.33 Similarly, the penetration indices for PMCl range from 11.0% to 12.7% whereas for YMCl and image file: d5ce00593k-t6.tif from 23.9% to 24.6%. Thus, As⋯Cl interactions in YMCl are stronger than those in PMCl and comparable to As⋯I secondary bonds in PNH4I. This is likely the result of differences between crystal structures of intercalates PMX, where M stands for alkali metal and ammonium cations, and X for halide anions, and compounds YMCl as well as very similar image file: d5ce00593k-t7.tif. Halide anions are involved in twelve As⋯X interactions in the former ones, while chloride anions only form six As⋯Cl secondary bonds in the latter compounds.
image file: d5ce00593k-f3.tif
Fig. 3 As⋯X bond valences (a) and penetration indices (b) plotted as a function of pressure for the studied intercalation compounds. Insets present Br (a) and Cl (b) anions interacting with twelve As atoms in PNH4Br and six As atoms YNH4Cl, respectively.

Penetration indices were chosen to compare the pressure evolution of As⋯X and As⋯O secondary bonds. They are plotted as a function of volume ratio V/V0, where V0 and V are the ambient pressure unit-cell volume and unit-cell volume at a given pressure, for the studied intercalates, as well as arsenic(III) oxide polymorphs, in Fig. 4. The ratio V/V0 was chosen as a free variable instead of pressure to account for different compressibility, or bulk moduli, of the crystal structures where secondary bonding is present. Such a choice of the coordinate system led to the observation of linear correlations between penetration indices and volume ratios. Again, the slopes of the linear correlations for the intercalates PNH4Br and PNH4I are equal within two standard uncertainties (−81(5) and −89.8(14), respectively) while the slope of −69.3(14) for intercalate YNH4Cl is slightly smaller in terms of absolute values, reflecting the fact that As⋯Cl secondary bond strengthening with pressure increase is less pronounced in YNH4Cl intercalates than the strengthening of secondary bonding in PNH4X intercalates. The analogous slope for the As⋯O secondary bonds in arsenolite is −114.1(5).


image file: d5ce00593k-f4.tif
Fig. 4 Penetration indices of As⋯X and As⋯O secondary bonds in the studied intercalates and As2O3 polymorphs plotted as a function of V/V0 unit-cell volume ratio. Data for arsenic(III) oxide polymorphs are taken from ref. 34–36.

Evolution of the arsenic coordination number and LEP stereoactivity with pressure

The first-order valence entropy coordination number of arsenic (1VECNAs) and arsenic LEP stereoactivity expressed as the magnitude of the resultant BVV vector on arsenic atoms are plotted in Fig. 5 as a function of the V/V0 ratio. The values of 1VECNAs for the studied intercalates change with the unit-cell volume ratio in a very similar manner to those of As2O3 polymorphs, and the observed changes are linear. 1VECNAs for the intercalates PNH4Br and YNH4Cl are almost the same even though the As⋯Cl secondary bonds are weaker than the As⋯Br secondary bonds.5 The reason for this is the same as in the case of penetration indices, namely the differences in crystal structures of the two types of intercalates. It is very likely, in our opinion, that chlorides tend to form the hydrated intercalates because the anions are relatively hard and are more keen on forming hydrogen bonds with water molecules than interactions with relatively soft LEPs of arsenic atoms. In the case of the softer bromide and iodide anions, no hydrated intercalates have been obtained so far as these anions are softer and form readily secondary interactions with arsenic LEPs.
image file: d5ce00593k-f5.tif
Fig. 5 First-order valence entropy coordination numbers of arsenic (a) and the magnitude of net BVV for arsenic (b) plotted as a function of the volume ratio V/V0 for the studied intercalates and for arsenic(III) oxide polymorphs. Data for arsenic(III) oxide polymorphs are taken from ref. 34–36.

The stereoactivity of arsenic LEPs in the studied intercalates does not change with pressure, like for arsenic(III) oxide, which indicates that pressure is an independent factor from LEPs causing strain in the coordination sphere of arsenic.27 For a given pressure, the LEP stereoactivity decreases slightly in the series YNH4Cl > PNH4Br > PNH4I which also testifies to the fact that secondary bond strength increases in the same order i.e. As⋯Cl < As⋯Br < As⋯I.

The correlation of the first-order valence entropy coordination number of arsenic with the magnitude of the resultant BVV vector on arsenic atoms for the studied intercalates, arsenic(III) oxide polymorphs under HP, and arsenic(III) oxycompounds deposited in the Inorganic Crystal Structure Database is plotted in Fig. 6. The grey markers corresponding to ambient-pressure structures follow a general correlation that LEP stereoactivity, quantified by the magnitude of arsenic BVV, decreases with increasing pressure. The points corresponding to the HP structures of the studied intercalates depart from the correlation, similarly to the HP structures of As2O3 polymorphs. While an increasing coordination number with increasing pressure is expected, the fact that LEP stereoactivity does not decrease at HP is surprising.37 Importantly, the points corresponding to the ambient-pressure structure of YNH4Cl and other hydrated intercalates YKCl and image file: d5ce00593k-t8.tif, represented by pentagons in Fig. 6, are located in the vicinity of the points corresponding to PMBr intercalates. This reinforces our suggestion that decreasing the number of Cl⋯As contacts in Y (and Y′) intercalates compensates for the fact that Cl anions are harder bases than Br anions and that As⋯Cl secondary bonds are of comparable strength to As⋯Br secondary bonds when chloride anions form six such bonds and bromide anions form twelve (see Fig. 3a with the insets and Fig. 6).


image file: d5ce00593k-f6.tif
Fig. 6 The magnitude of net BVV for arsenic plotted as a function of the first-order valence entropy coordination number of arsenic for the studied intercalates, As2O3 polymorphs, and arsenic(III) oxycompounds deposited in the Inorganic Crystal Structure Database. HP in the legend denotes high-pressure structures. The larger markers indicated at the bottom right correspond to ambient-pressure structures of intercalates. The black arrow indicates the compression of the systems. Data for arsenic(III) oxide polymorphs, for arsenic(III) oxycompounds, and for intercalates are taken from ref. 5, 9, 27 and 34–36.

Conclusions

Crystal structures of three arsenic(III) oxide intercalates with ammonium halides have been determined under high pressure for the first time. They do not undergo any phase transitions up to the highest pressures reached in this study (12 GPa for YNH4Cl, 15 GPa for PNH4Br and 11 GPa for PNH4I). Analysis of the pressure dependence of penetration indices of As⋯X secondary bonds, where X stands for a halogen, revealed that their dependence becomes linear when pressure is replaced by the unit-cell volume ratio V/V0, where V0 is the ambient pressure volume. This is caused by the fact that the V/V0 ratio takes into account different compressibility, or bulk moduli, of the compared crystal structures. Not only the penetration indices but also the first-order valence entropy coordination number of arsenic decrease linearly with volume ratio V/V0. Both these relationships are also linear for As⋯O secondary bonds present in arsenic(III) oxide polymorphs. The stereoactivity of arsenic LEPs is virtually unchanged upon pressure increase indicating that pressure increases the arsenic coordination number by making the crystal structures “more crowded” without decreasing the LEP stereoactivity similar to arsenic(III) oxide polymorphs. It will be very interesting to carry out an analogous study for a series of As2O3 intercalates with potassium halides, which includes anhydrous PKCl, PKBr, and PKI and hydrated YKCl. It should allow us to verify our suggestion that the increased strength of As⋯Cl secondary bonds in intercalate YNH4Cl compared to the expectations from the trend of decreasing strength of secondary interactions As⋯I > As⋯Br > As⋯Cl is the result of a smaller number of secondary bonds formed by chloride anions in intercalate YNH4Cl than in compounds PNH4X.

Author contributions

P. A. G. – conceptualization, investigation, writing – original draft, writing – review & editing; S. M. – investigation, writing – review & editing; M. E. – investigation, writing – review & editing; F. A. – resources, methodology, writing – review & editing; R. M. – resources, writing – review & editing; K. F. D. – investigation, formal analysis, writing – review & editing.

Conflicts of interest

There are no conflicts to declare.

Data availability

Supplementary information available: figures with the studied crystals' mosaicities and unit cell parameters plotted as a function of pressure; a table with linear bulk moduli. See DOI: https://doi.org/10.1039/D5CE00593K.

CSD 2440780–2440816 contain the supplementary crystallographic data for this paper.

Raw diffraction data are available at Zenodo at https://doi.org/10.5281/zenodo.15356607.

Acknowledgements

This research has been funded by the Polish National Science Centre (project no. 2020/39/D/ST4/00128). Elettra Sincrotrone Trieste is gratefully acknowledged for providing beam time for the experiments at the Xpress beamline (Proposals 20230005 and 20240142). Fruitful discussions with Ross J. Angel and G. Diego Gatta are gratefully acknowledged.

Notes and references

  1. T. I. U. of P. and A. Chemistry (IUPAC), IUPAC – intercalation compounds (I03076), https://goldbook.iupac.org/terms/view/I03076, (accessed May 22, 2024).
  2. R. A. Klemm, Phys. C, 2015, 514, 86–94 CrossRef CAS .
  3. A. Krzton-Maziopa, Z. Shermadini, E. Pomjakushina, V. Pomjakushin, M. Bendele, A. Amato, R. Khasanov, H. Luetkens and K. Conder, J. Phys.: Condens. Matter, 2011, 23, 052203 CrossRef CAS PubMed .
  4. W. Zeng, Y. Tian, H. Zeng, Z. Lin and G. Zou, Angew. Chem., Int. Ed., 2025, 64, e202422818 CrossRef CAS PubMed .
  5. M. Dąbrowski, W. Wrześniewska, A. Prystupiuk, J. Zachara and P. A. Guńka, Cryst. Growth Des., 2023, 23, 6081–6085 CrossRef .
  6. F. Pertlik, J. Solid State Chem., 1987, 70, 225–228 CrossRef CAS .
  7. P. A. Guńka, K. Kraszewski, Y.-S. Chen and J. Zachara, Dalton Trans., 2014, 43, 12776–12783 RSC .
  8. M. A. Domański, K. Kraszewski, P. Paluch and P. A. Guńka, Cryst. Growth Des., 2021, 21, 5215–5222 CrossRef .
  9. P. Michalak, P. Paluch and P. A. Guńka, Cryst. Growth Des., 2022, 22, 711–717 CrossRef CAS .
  10. M. Edstrand and G. Blomqvist, Ark. Kemi, 1955, 8, 245–256 CAS .
  11. W. Wrześniewska, P. Paluch and P. A. Guńka, Acta Crystallogr., Sect. B: Struct. Sci., Cryst. Eng. Mater., 2023, 79, 207–212 CrossRef PubMed .
  12. A. Dewaele, M. Torrent, P. Loubeyre and M. Mezouar, Phys. Rev. B: Condens. Matter Mater. Phys., 2008, 78, 104102 CrossRef .
  13. P. Lotti, S. Milani, M. Merlini, B. Joseph, F. Alabarse and A. Lausi, J. Synchrotron Radiat., 2020, 27, 222–229 CrossRef CAS .
  14. CrysAlisPro Software system ver. 171.43 Rigaku Oxford Diffraction, Oxford, UK, 2023 Search PubMed .
  15. G. M. Sheldrick, Acta Crystallogr., Sect. C: Struct. Chem., 2015, 71, 3–8 Search PubMed .
  16. O. V. Dolomanov, L. J. Bourhis, R. J. Gildea, J. A. K. Howard and H. Puschmann, J. Appl. Crystallogr., 2009, 42, 339–341 CrossRef CAS .
  17. A. L. Spek, J. Appl. Crystallogr., 2003, 36, 7–13 CrossRef CAS .
  18. F. Birch, Phys. Rev., 1947, 71, 809–824 CrossRef CAS .
  19. J. Gonzalez-Platas, M. Alvaro, F. Nestola and R. Angel, J. Appl. Crystallogr., 2016, 49, 1377–1382 CrossRef CAS .
  20. I. D. Brown, The Chemical Bond in Inorganic Chemistry: The Bond Valence Model, Oxford University Press, Oxford, 2016 Search PubMed .
  21. O. C. Gagné and F. C. Hawthorne, Acta Crystallogr., Sect. B: Struct. Sci., Cryst. Eng. Mater., 2015, 71, 562–578 CrossRef PubMed .
  22. N. E. Brese and M. O'Keeffe, Acta Crystallogr., Sect. B: Struct. Sci., 1991, 47, 192–197 CrossRef .
  23. I. D. Brown, Bond-valence parameters, https://www.iucr.org/resources/data/datasets/bond-valence-parameters, (accessed April 25, 2023).
  24. I. D. Brown, P. Klages and A. Skowron, Acta Crystallogr., Sect. B: Struct. Sci., 2003, 59, 439–448 CrossRef PubMed .
  25. J. Zachara, Inorg. Chem., 2007, 46, 9760–9767 CrossRef CAS .
  26. J. Echeverría and S. Alvarez, Chem. Sci., 2023, 14, 11647–11688 RSC .
  27. P. A. Guńka and J. Zachara, Acta Crystallogr., Sect. B: Struct. Sci., Cryst. Eng. Mater., 2019, 75, 86–96 CrossRef .
  28. G. Kresse and J. Furthmüller, Phys. Rev. B: Condens. Matter Mater. Phys., 1996, 54, 11169–11186 CrossRef CAS PubMed .
  29. H. J. Monkhorst and J. D. Pack, Phys. Rev. B: Solid State, 1976, 13, 5188–5192 CrossRef .
  30. G. Kresse and D. Joubert, Phys. Rev. B: Condens. Matter Mater. Phys., 1999, 59, 1758–1775 CrossRef CAS .
  31. S. Grimme, J. Antony, S. Ehrlich and H. Krieg, J. Chem. Phys., 2010, 132, 154104 CrossRef PubMed .
  32. S. Grimme, S. Ehrlich and L. Goerigk, J. Comput. Chem., 2011, 32, 1456–1465 CrossRef CAS PubMed .
  33. Both Y and Y′ intercalates are hydrated. The only difference between them lies in the structure of the disordered layer located at z = 1/2.
  34. P. A. Guńka, K. F. Dziubek, A. Gładysiak, M. Dranka, J. Piechota, M. Hanfland, A. Katrusiak and J. Zachara, Cryst. Growth Des., 2015, 15, 3740–3745 CrossRef .
  35. P. A. Guńka, M. Dranka, M. Hanfland, K. F. Dziubek, A. Katrusiak and J. Zachara, Cryst. Growth Des., 2015, 15, 3950–3954 CrossRef .
  36. P. A. Guńka, M. Hanfland, Y.-S. Chen and J. Zachara, CrystEngComm, 2021, 23, 638–644 RSC .
  37. W. Grochala, R. Hoffmann, J. Feng and N. W. Ashcroft, Angew. Chem., Int. Ed., 2007, 46, 3620–3642 CrossRef CAS .

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.