ACS Publications. Most Trusted. Most Cited. Most Read
Mechanochemical Synthesis and Magnetic Properties of the Mixed-Valent Binary Silver(I,II) Fluorides, AgI2AgIIF4 and AgIAgIIF3
My Activity

Figure 1Loading Img
  • Open Access
Article

Mechanochemical Synthesis and Magnetic Properties of the Mixed-Valent Binary Silver(I,II) Fluorides, AgI2AgIIF4 and AgIAgIIF3
Click to copy article linkArticle link copied!

Open PDFSupporting Information (1)

Journal of the American Chemical Society

Cite this: J. Am. Chem. Soc. 2024, 146, 44, 30510–30517
Click to copy citationCitation copied!
https://doi.org/10.1021/jacs.4c11772
Published October 24, 2024

Copyright © 2024 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY 4.0 .

Abstract

Click to copy section linkSection link copied!

Fluoridoargentates(II) represent a fascinating class of silver(II) compounds with structural and magnetic similarities to cuprate superconductors. However, their synthesis is challenging, leaving their properties largely underexplored and hindering the discovery of new phases. This study introduces mechanochemistry as a novel approach for the synthesis of fluoridoargentates(II), avoiding the use of anhydrous HF or elemental fluorine and employing readily available equipment. Notably, ball milling of commercially available precursors successfully produced the long-sought-after first two examples of binary mixed-valent silver(I,II) phases, AgI2AgIIF4 (Ag3F4) and AgIAgIIF3 (Ag2F3). While the AgI2AgIIF4 phase was obtained at room temperature, the AgIAgIIF3 phase is metastable and required milling under cryogenic conditions. Characterization by synchrotron powder X-ray and neutron diffraction revealed that AgI2AgIIF4 crystallizes in the P21/c space group and is isostructural to β-K2AgF4. In this crystal structure, [AgIIF2F4/2]2– distorted octahedral units with 4 + 2 coordination, extend parallel to a-crystallographic axis giving a quasi-one-dimensional canted antiferromagnetic character, as shown by magnetic susceptibility. The triclinic perovskite AgIAgIIF3 phase adopts the P1̅ space group, is isostructural to AgCuF3 and also shows features of a one-dimensional antiferromagnet. This mechanochemical approach, also successfully applied to synthesize β-K2AgF4, is expected to expand the field of silver(II) chemistry, accelerating the search for silver analogs to cuprate superconductors and potentially extending to other cations in unusual oxidation states.

This publication is licensed under

CC-BY 4.0 .
  • cc licence
  • by licence
Copyright © 2024 The Authors. Published by American Chemical Society

Introduction

Click to copy section linkSection link copied!

Systems containing S = 1/2 cations offer a rich playground to explore quantum effects such as spin liquids, spin-Peierls instabilities, superconductivity and the like. (1,2) Compared to other metal cations like Cu2+, Ti3+, or V4+, systems containing Ag2+ are less explored although they show great promise for very intriguing physics such as exotic magnetism and even potential superconductivity as cuprate analogues. (3,4) However, unlike Cu2+, which is in a common oxidation state, Ag2+ is an unstable and very reactive species. (5,6) In oxides, Ag2+ disproportionates to Ag+ and Ag3+. Fluorine is the only element that is sufficiently electronegative to readily stabilize Ag in a +2 oxidation state. These compounds, fluoridoargentates(II), are truly exceptional in various aspects. (7) Even the most fundamental binary compound, AgF2, displays intriguing structural and electronic properties similar to the parent compound of the cuprate superconductors, La2CuO4, (8) which features strong mixing of Ag(4d9) and F(2p2) orbitals, a significant covalency of the Ag–F bond, (9) and significant magnetic superexchange in two dimensions. (10) Similar to Cu2+(3d9), the Ag2+(4d9), S = 1/2, can also be found in a Jahn–Teller elongated or compressed octahedral environment.
By analogy with cuprates, where superconductivity can be achieved by charge doping, mixed-valence silver(I,II) or silver(II,III) fluorides are very enticing as potential candidates for metallization analogous to charge-doped cuprates. (3) However, synthesizing these phases has proven challenging, with silver(I,II) fluorides being particularly elusive. Currently, the binary fluorine–silver system only comprises six phases: Ag2F, (11−13) AgF, (14,15) AgF2, (16−18) AgF3, (19,20) Ag2F5, (21) and Ag3F8. (22) In these compounds, silver adopts the +1/2, +1, +2 and +3 oxidation states. Interestingly, only two binary mixed-valence silver fluorides are known up to date: Ag2F5 or AgIIF[AgIIIF4], and Ag3F8 or AgII[AgIIIF4]2. In both cases, silver is in a mixed +2/+3 oxidation state. No binary Ag–F phases exist with silver in a mixed +1/+2 oxidation state. However, theoretical calculations have been performed to search for the most stable compounds and structures in the binary Ag–F system. (23−25) Of outmost interest are mixed valence silver(I,II) binary fluorides, highly sought-after phases as they are believed to be the charged-doped AgF2 equivalents. These theoretical studies suggested that two novel stoichiometries with silver in a mixed +1/+2 oxidation state, namely Ag3F4 and Ag2F3, could be thermodynamically stable at ambient conditions. The hypothetical AgI2AgIIF4 or Ag3F4 phase was predicted (23) to adopt a monoclinic P21/c Na2CuF4-type (26) structure analogous to β-K2AgF4. (27) A theoretical investigation of silver fluorides under pressure proposed a tetragonal I4̅2d CdMn2O4-type structure as the ground state structure of AgI2AgIIF4 which was predicted to be stable up to 19 GPa. (24) The existence of another silver(I,II) fluoride with the formula AgIAgIIF3 or Ag2F3 was also predicted by density functional theory calculations. Three nearly energy-equivalent structure types were predicted: P1̅ (NaCuF3-type), Pbnm (KAgF3-type) and, Pmcn (CuTeO3-type structure). (23) Another theoretical study proposed a structure with the P21/m space group (distorted CaIrO3-type structure). (24) The calculated energies of formation for the most stable polytype was −0.09 eV per formula unit for AgIAgIIF3 and −0.21 eV for AgI2AgIIF4. (23) The more exothermic energy of formation for AgI2AgIIF4 compared to AgIAgIIF3 could indicate that AgI2AgIIF4 is the most thermodynamically stable product of the reaction between AgF2 and AgF. Indeed, Ag2F3 was predicted to be only marginally stable with respect to decomposition into Ag3F4 and AgF2 at ambient pressure. (24)
Following these predictions, experimental efforts to synthesize these phases have fallen short, as current synthetic methods have not been able to provide access to these compounds which could be metastable. The main synthetic challenges were the poor solubility of the main precursor, AgF2, in anhydrous HF and thermal decomposition of AgF2 at temperatures <400 °C. Mechanochemistry could overcome the disadvantages of conventional synthesis methods, by offering room-temperature reaction conditions, enhanced kinetics, and solvent-free synthesis. (28−31) Mechanochemistry represents a paradigm shift in the synthesis of materials, enabling the formation of metastable phases that are difficult to obtain using high-temperature methods. While there are numerous reports on the high-energy milling of oxides, (32) indicating that it is an established synthetic technique, the mechanochemistry of fluorides on the other hand is still in its early stages of development, (33,34) with the mechanochemistry of fluoridoargentates(II) being virtually an unexplored area of research at the onset of this work.
In this study, it is demonstrated that mechanochemistry is an effective approach to the formation of the anticipated mixed-valent binary silver(I,II) fluorides and showcases that it could be applied to the synthesis of fluoridoargentates(II) and other compounds with elements in unusual and mixed-valent oxidation states.

Results and Discussion

Click to copy section linkSection link copied!

AgI2AgIIF4

For the synthesis of AgI2AgIIF4, a 2:1 AgF–AgF2 molar ratio was ball milled at room temperature. Experimental details are provided in the Supporting Information. A laboratory powder X-ray diffraction (PXRD) (λ = 0.5609 Å) analysis of the resulting pale light-brown polycrystalline sample confirmed the formation of the new Ag3F4 phase [88.7(2) wt %], as well as the presence of AgF [11.3(6) wt %] as a result of decomposition of AgF2 during milling (Figure 1a). After the successful mechanochemical synthesis, an attempt was made to synthesize Ag3F4 via a solid-state route, using a method similar to that employed for M2AgF4 (M = Na–Cs) compounds. (35) However, to minimize the decomposition of AgF2, the optimized procedure involved a lower synthesis temperature (300 °C), which required longer dwell times of up to 3 days. A comparison of the laboratory diffractograms of the ball-milled and solid-state synthesized samples (Figure S1) reveals that, while both samples exhibit diffraction peaks belonging to the new compound, the solid-state sample contains more impurities. On the other side, the ball-milled sample shows significant peak broadening. This broadening suggests the presence of nanometer-sized crystallites and defects or microstrain induced during the milling process. A higher background is also observed in the milled sample, indicating the presence of an amorphous phase. Similar observations were also noted in the milling of oxides. (32) Although the solid-state sample appeared to contain a larger amount of impurities than the ball-milled one, high-resolution synchrotron PXRD data (λ = 0.3497 Å) was collected at room temperature on the sample prepared by the solid-state method for indexation and structure refinement, due to its higher crystallinity (Figure 1b). Unassigned peaks that did not belong to unreacted reagents (AgF or AgF2) could be indexed to a monoclinic unit cell in the P21/c space group (Figure S2) with unit cell parameters similar to Na2CuF4. (26)

Figure 1

Figure 1. (a) Rietveld refinement of the laboratory PXRD data (Ag Kα radiation, λ = 0.5609 Å) of mechanochemically synthesized sample of AgI2AgIIF4 collected at room temperature. Quantitative analysis: AgI2AgIIF4: 88.7(2) wt %; AgF: 11.3(6) wt %. (b) Three-phase Rietveld refinement of the synchrotron data collected on the AgI2AgIIF4 sample prepared by solid-state method. Quantitative analysis: AgI2AgIIF4: 76.2(1) wt %; AgF: 22.9(1) wt %; AgF2: 0.9(5) wt %. (c) Unit cell of the AgI2AgIIF4 (Ag3F4) crystal structure and a chain of edge-sharing [AgIIF2F4/2]2– distorted octahedral units with 4 + 2 coordination, which extends parallel to a-crystallographic axis. (d) The 5 K NPD pattern of the ball-milled AgI2AgIIF4 sample showing that the only impurity is small amounts of diamagnetic AgF. Quantitative analysis: AgI2AgIIF4: 89.4(1) wt %; AgF: 10.6(1) wt %.

A Rietveld refinement analysis performed using the Na2CuF4-structure type with a monoclinic P21/c unit cell resulted in an excellent fit (Figure 1b) and the following unit cell parameters: a = 3.56910(8) Å, b = 9.8866(2) Å, c = 5.9829(2) Å, β = 92.832(2)°, V = 210.857(4) Å3, Z = 2 (Table S1a, Figures 1c and S3). The observed AgI–F distances [2.391(7)–2.884(8) Å] and three sets of AgII–F bond lengths [2.110(6), 2.119(7) and 2.529(7) Å] are in the expected range (Table S1b). (15,18,27,36) The edge-sharing distorted octahedral [AgF6]4– units, which display axially elongated 4 + 2 coordination (Figures 1c and S4), are connected into chains extending in a-crystallographic direction. The bond valence sum (BVS) (37) analysis confirms the mixed +1/+2 valence state of silver (Table S1b). Furthermore, the monoclinic structural model of AgI2AgIIF4 was also confirmed by the Rietveld refinement of neutron powder diffraction (NPD) (λ = 2.41 Å) data collected at 5 K on the mechanochemically synthesized AgI2AgIIF4 sample (Figure 1d). Taking into account the difference in measurement temperatures, NPD-derived unit cell parameters are in excellent agreement with the synchrotron PXRD results, giving a = 3.5528(4) Å, b = 9.7931(12) Å, c = 5.9406(8) Å, β = 92.614(3)°, V = 206.48(8) Å3 (Table S2a). The AgI–F bond distances [2.373(2)–2.834(3) Å] and three sets of AgII–F lengths [2.096(2), 2.080(2) and 2.526(2) Å] (Table S2b) are also in good agreement with the structure determined by PXRD. Moreover, the amount of AgF impurity [10.6(1) wt %] detected in the NPD data is in accordance with the PXRD results [11.3(6) wt %] (Figure 1a).
The Rietveld refinement analysis of the synchrotron PXRD data measured on the sample synthesized by solid-state method also revealed a deviation from the expected weight ratio of 1.74:1 for AgF to AgF2 if the sample decomposed to the starting materials. Instead, it showed 22.9(1) wt % of AgF and only 0.9(5) wt % of AgF2 (Figure 1b). It can be assumed that this larger amount of AgF is the result of the sample decomposition in the capillary, as indicated by the loss of transparency of the glass. When the PXRD measurement was performed on a fresher sample, it showed a lower AgF content of 14.3(1) wt % (Figure S5). Nevertheless, mechanochemically synthesized samples of Ag3F4 were of superior purity than samples prepared by classical solid-state approach [88.7(2) and 89.4(1) vs 85.4(2) wt %].
An experimental indication of the existence of AgI2AgIIF4 phase was also obtained from Raman measurements performed on AgF2 using a green excitation laser (λ = 532 nm). Longer illumination times led to changes in the Raman spectrum of AgF2; a decrease of the main band at 260 cm–1 and a significant increase of the minor band at 420 cm–1 (Figure S6) indicated a possible photochemical decomposition. Indeed, this spectrum is in excellent agreement with the Raman spectrum of the mechanochemically synthesized Ag3F4 sample. Both samples share the same spectral features, however the peaks in the mechanochemical sample are much broader. A similar observation on laser decomposition of AgF2 has been previously made but without the confirmation of the Ag3F4 phase formation. (38) Since AgI2AgIIF4 is isostructural to β-K2AgF4, (27) similarities in their Raman spectra can be expected. To enable a direct comparison of Raman spectra measured under the same conditions, β-K2AgF4 was also synthesized mechanochemically (Figure S7). The successful formation of this phase further demonstrates the effectiveness of mechanochemistry for the synthesis of fluoridoargentates(II). The Raman spectrum of AgI2AgIIF4 (Figure 2a) displays the most intense peak at 422 cm–1, which could be attributed to the symmetric vibrations of the [AgF4]2– subunits and a shoulder at 488 cm–1. This is comparable to the peaks (423 and 489 cm–1) observed in the spectrum of β-K2AgF4, which is in good accordance with the literature reported values. (27) Overall, the Raman spectra of both compounds display the same spectral features, confirming the structural similarities between the two phases. The results of the vibrational spectra thus also confirm that AgI2AgIIF4 is isostructural to β-K2AgF4.

Figure 2

Figure 2. (a) Raman spectra of AgI2AgIIF4 and β-K2AgF4 synthesized by ball milling. (b) Temperature-dependent magnetic susceptibility of mechanochemically prepared AgI2AgIIF4 sample measured in a magnetic field of 1 kOe. The inset shows the Curie–Weiss fit to the inverse of magnetic susceptibility in the 15–300 K range. (c) Zero field cooled (ZFC) and field cooled (FC) data in the 2–25 K temperature region showing the upturn in the magnetic susceptibility characteristic of AFM, while the Bonner–Fisher fit for the 2–15 K region is shown in the inset. (d) Magnetization of AgI2AgIIF4 as a function of magnetic field measured at 2 and 5 K. The hysteresis shown in the inset indicates a FM ground state.

Given the structural similarities, the magnetic properties of AgI2AgIIF4 can also be expected to resemble those of the β-K2AgF4 phase. (27) The temperature-dependent magnetic susceptibility data (Figure 2b) shows that AgI2AgIIF4 behaves like a paramagnet from room temperature down to about 5 K. At about 4 K, a broad peak can be observed, indicative of low-dimensional antiferromagnetic (AFM) properties, after which an abrupt increase in the magnetic susceptibility with lowering the temperature down to 3 K is observed, indicating a ferromagnetic (FM) transition. The value of the FM transition temperature was determined as 2.8(1) K from the maximum of the first derivative of the magnetic susceptibility, dχ/dT. The ball-milled and solid-state synthesized samples exhibit nearly identical magnetic responses (Figure S8), although the ball-milled sample displays a more pronounced broad maximum (see arrow).
The inverse magnetic susceptibility data was corrected for a constant diamagnetic contribution and fitted from 15 to 300 K with the Curie–Weiss law (Figure 2b, inset). The derived Curie constant of 0.48(1) emu·K/mol is in good agreement with the expected value for the S = 1/2 Ag2+ cations. A negative value of the Curie–Weiss temperature was determined, θCW = −3.6(2) K, suggesting that the dominant exchange is AFM. The 2–25 K region (Figure 2c) clearly shows the presence of a broad peak before the FM transition with a maximum at about 4 K, which is characteristic of a 1D AFM. A fit to the region around TN (2–15 K) with a uniform AFM spin-1/2 chain Bonner–Fisher model (39) (Figure 2c, inset) is excellent and the obtained parameters, g = 2.3 and |J1D|/kB = 2.9 K are in good agreement with the values expected for a 1D AFM model, further supporting the 1D spin correlations above the TN for the AgI2AgIIF4 phase.
To gain a deeper insight into the magnetism of AgI2AgIIF4, the M(H) curves were measured (Figure 2d). At 5 K, the magnetization shows a linear relationship with the applied magnetic field, without any remnant magnetization or coercivity. However, at 2 K, which is below the transition temperature, a pronounced hysteresis loop is observed, characterized by a remnant magnetization of approximately 0.0165 μB/f.u. and a coercive field of Hc = 0.076 kOe. Even at the largest applied magnetic field of 50 kOe, the magnetization value remains modest at 0.07 μB/f.u. and increases linearly with the strength of the magnetic field. The magnetization value falls significantly below the anticipated saturation magnetization of ≈1 μB/f.u. for the S = 1/2 spin of the Ag2+ ion, which could be a result of a canted antiferromagnetism. A rough estimate of the canting angle (∼0.9°) in zero magnetic field could be obtained by comparing the measured remnant magnetization to the full magnetic moment of 1 μB per Ag2+ ion. The NPD measurements performed above (5 K) and below (1.5 K) the transition temperature [2.8(1) K] observed in the magnetic susceptibility data of the ball-milled AgI2AgIIF4 sample show only one weak feature at around 16° 2θ in the 1.5 K data set (Figure S9, inset). This is not surprising, as weak magnetic Bragg peaks are in fact expected due to the low-dimensional magnetic structure and the small moment expected from the Ag2+ (S = 1/2). However, this sole peak was not sufficient to determine the magnetic structure of AgI2AgIIF4.

AgIAgIIF3

The second stoichiometry that was thoroughly investigated in this study was a 1:1 AgF–AgF2 molar ratio. The experimental details are described in the Supporting Information. Note that cooling the milling jars with liquid nitrogen was essential in stabilizing this new phase and minimizing its decomposition to AgI2AgIIF4 and AgF2. The laboratory PXRD data on the light-brown powder (Figure 3a) confirmed the successful mechanochemical formation of a new phase AgIAgIIF3 [95.6(1) wt %] with minor impurities Ag3F4 [3.3(2) wt %] and AgF2 [1.1(3) wt %]. Solid state synthesis was also performed at 300 °C and the synchrotron PXRD data collected on this sample was used to index the peaks of the new AgIAgIIF3 phase (Figure S10). A triclinic P1̅ space group was found, with a unit cell similar to that of AgCuF3, (40) resembling a distorted MAgF3 structure (M = K, Rb and Cs). (41) A Rietveld refinement performed using the triclinic P1̅ space group resulted in excellent fit and the following unit cell parameters: a = 5.9577(1) Å, b = 5.8217(1) Å, c = 8.5467(2) Å, α = 91.565(2), β = 90.644(2), γ = 85.977(1)°, V = 295.582(6) Å3, Z = 4 (Figure S11, Table S3a). Thus, AgIAgIIF3 is isostructural to AgCuF3, (40) and adopts the NaCuF3 structure type (Figure 3b). (42) In the crystal structure of Ag2F3, there are two nonequivalent silver(I) atoms, both coordinated with seven fluorine atoms [2.341(16)–3.072(20) Å]. Moreover, four nonequivalent silver(II) atoms are all coordinated with six fluorine atoms, forming a [AgF6]4– octahedra with four short [2.066(11)–2.161(13) Å] and two long [2.403(11)–2.441(16) Å] AgII–F distances (Table S3b, Figures S12 and S13). The octahedra thus exhibit an axial elongation (4 + 2) and could therefore be Jahn–Teller active. The corner-sharing [AgF6/2] units are connected into a distorted perovskite structure, resembling a structural arrangement of the KAgF3 compound (Figure S12). (41,43) The BVS analysis (37) (Table S3b) confirms the valence state of the silver(I) and silver(II) atoms and provides further support for the correctness of the determined crystal structure.

Figure 3

Figure 3. (a) Rietveld refinement of the laboratory PXRD data (Ag Kα radiation, λ = 0.5609 Å) of mechanochemically synthesized sample of AgIAgIIF3, measured at room temperature. Quantitative analysis: AgIAgIIF3: 95.6(1) wt %; AgI2AgIIF4: 3.3(2) wt %; AgF2: 1.1(3) wt %. (b) Unit cell of the AgIAgIIF3 (Ag2F3) crystal structure with apex-sharing [AgIIF6/2] distorted octahedral units with 4 + 2 coordination. (c) Raman spectrum of ball-milled AgIAgIIF3. For comparison, the Raman spectrum of KAgF3 is also shown. (d) Weight fraction of phases in the ball-milled samples obtained from Rietveld refinement of the laboratory PXRD data as a function of milling time. The amorphous phase was not accounted in the refinement; the errors are smaller than the symbols in the graph.

The Raman spectrum of AgIAgIIF3 was measured and compared to the literature reported one for KAgF3. (44) For a detailed comparison of Raman spectra measured under the same conditions, KAgF3 was also synthesized. Raman spectra of Ag2F3 and KAgF3 exhibit very similar features (Figure 3c) with the most intense peaks observed at 480, 378 and 467, 380 cm–1, respectively.
It is noteworthy, that the most intense Raman active mode of AgI2AgIIF4 located at 422 cm–1 is also present in the Raman spectrum of AgIAgIIF3, albeit with a very weak signal due to the low content. The presence of the Ag3F4 phase could be explained by the observation that the mechanochemically synthesized AgIAgIIF3 sample stored in the glovebox at room temperature decomposes into AgF2 and AgI2AgIIF4 over time (Figure S14).
To gain a deeper insight into the reaction mechanism of Ag3F4 formation, time-dependent milling experiments with sampling every 15 min were performed (Figure S15). The analysis reveals that after 15 min of milling, the majority of the AgF was consumed and the milled sample consisted mainly of AgF2, AgIAgIIF3 and AgI2AgIIF4 (Figure 3d). After 60 min of milling, the sample contained AgIAgIIF3 with traces of the AgI2AgIIF4 phase. It can thus be assumed that AgIAgIIF3 is formed from the reaction of the intermediate AgI2AgIIF4 phase with unreacted AgF2 (eq 1)
Ag2IAgIIF4+AgF22AgIAgIIF3
(1)
Furthermore, the magnetic susceptibility data measured on several mechanochemically synthesized samples reveals intriguing features that provide valuable insights into the magnetic properties of AgIAgIIF3 (Figures 4a and S16).

Figure 4

Figure 4. Temperature-dependent magnetic susceptibility, χ, of AgIAgIIF3, measured in a magnetic field of (a) 1 and (b) 10 kOe. The inset in (a) shows the first derivative indicating a deviation from the Curie–Weiss law at 7.3(7) K. The inset in (b) shows the inverse of the magnetic susceptibility (FC) measured at 10 kOe. (c) Plot of the effective magnetic moment, μeff, as a function of temperature for FC and ZFC curves at 1 kOe and for the FC curve measured at 10 kOe. (d) Magnetization of AgIAgIIF3 as a function of magnetic field, measured at 2 K.

A weak AgF2 impurity signal is observed at approximately 164 K and quantified from the χ·T data to about 0.8(3) wt % AgF2, which is in good agreement with the value of 1.1(3) wt % obtained from the PXRD analysis (Figure 3a). The presence of AgF2 is expected due to either the 10 mol % excess of AgF2 used in the synthesis or a decomposition of the Ag2F3 phase into Ag3F4 phase and AgF2. A weak FM signal corresponding to the Ag3F4 phase [3.3(2) wt %] is also observed. It should be noted that when the magnetization is measured on the AgIAgIIF3 sample synthesized mechanochemically with a stoichiometric 1:1 AgF–AgF2 molar ratio, a smaller amount of AgF2 is present, but a larger amount of the Ag3F4 phase is observed as a secondary phase, which is minimized when an excess of AgF2 is used in the synthesis.
The first derivative of magnetic susceptibility indicates the temperature of 7.3(7) K at which the deviation from the Curie–Weiss law can be observed (Figure 4a, inset). The magnetic susceptibility was also measured in a ten times larger magnetic field of 10 kOe to minimize the influence of the FM signal of AgF2 and other possible FM impurities (Figure 4b). The high-temperature range (200–300 K) of the 10 kOe measurement shows a persistent, continuous decrease in susceptibility with decreasing temperature, indicative of AFM interactions persisting even at elevated temperatures. The observed behavior resembles KAgF3, a 1D antiferromagnet. (44) This hypothesis is also supported by the fact that the effective magnetic moment at 300 K is well below the theoretically expected value of 1.7 μB, reaching only 0.8 μB (Figure 4c). The effective magnetic moment also decreases with decreasing temperature and reaches a value of approximately 0.15 μB, indicating AFM interactions are predominant throughout the whole measured temperature range. In addition, the magnetization measurements (Figure 4d) show that the magnetic moments do not attain saturation even under the influence of high magnetic fields, a characteristic that strongly suggests the presence of AFM interactions.

Conclusions

Click to copy section linkSection link copied!

In this study, mechanochemistry was employed as a novel synthetic method for fluoridoargentates(II) and provided the first experimental evidence for two binary mixed-valent silver(I,II) fluorides, AgI2AgIIF4 (Ag3F4) and AgIAgIIF3 (Ag2F3). The AgI2AgIIF4 phase is stable at room temperature, whereas the AgIAgIIF3 phase appears to be metastable and required milling at liquid nitrogen temperatures. Synchrotron PXRD, NPD, and Raman spectroscopy studies confirmed the Na2CuF4-type structure of AgI2AgIIF4, which is similar to that of β-K2AgF4. This novel compound shows canted antiferromagnetism below 2.8(1) K. On the other side, the triclinic perovskite AgIAgIIF3 phase, isostructural to AgCuF3, shares structural features with KAgF3, and shows characteristics of a one-dimensional antiferromagnet. This work thus demonstrates the effectiveness of mechanochemical approach for the synthesis of fluoridoargentates(II), as it provides samples of Ag3F4, Ag2F3, and β-K2AgF4 with superior purity in comparison to the conventional solid-state thermal route. Moreover, mechanochemistry enables a low-temperature synthesis, addressing the challenges posed by Ag2+ reactivity and thermal instability of AgF2. This approach not only facilitates the discovery of new silver(II) compounds but also enables the expansion of this relatively underexplored area of chemistry.

Supporting Information

Click to copy section linkSection link copied!

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/jacs.4c11772.

  • The experimental procedures, description of the characterization techniques, and results of analyses (PDF)

Accession Codes

Deposition numbers 2321112 (for Ag2F3) and 2321184 (for Ag3F4) contain the supplementary crystallographic data for this paper. These data are provided free of charge by the joint Cambridge Crystallographic Data Centre and Fachinformationszentrum Karlsruhe Access Structures service.

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Author
  • Authors
    • Matic Belak Vivod - Jožef Stefan Institute, Jamova cesta 39, 1000 Ljubljana, SloveniaJožef Stefan International Postgraduate School, Jamova cesta 39, 1000 Ljubljana, Slovenia
    • Zvonko Jagličić - Institute of Mathematics, Physics and Mechanics, 1000 Ljubljana, SloveniaFaculty of Civil and Geodetic Engineering, University of Ljubljana, Jamova cesta 2, 1000 Ljubljana, Slovenia
    • Graham King - Canadian Light Source, 44 Innovation Blvd, Saskatoon, S7N 2V3 Saskatchewan, CanadaOrcidhttps://orcid.org/0000-0003-1886-7254
    • Thomas C. Hansen - Institut Laue-Langevin, 38042 Grenoble Cedex 9, FranceOrcidhttps://orcid.org/0000-0003-4611-2393
    • Matic Lozinšek - Jožef Stefan Institute, Jamova cesta 39, 1000 Ljubljana, SloveniaJožef Stefan International Postgraduate School, Jamova cesta 39, 1000 Ljubljana, SloveniaOrcidhttps://orcid.org/0000-0002-1864-4248
  • Notes
    The authors declare no competing financial interest.

Acknowledgments

Click to copy section linkSection link copied!

Support from the Slovenian Research and Innovation Agency (J2-2496, P2-0105, P2-0348); the Marie Skłodowska-Curie Individual Fellowship (101031415) and the European Research Council Starting Grant (950625) under the European Union’s Horizon 2020 Research and Innovation Programme; and the Jožef Stefan Institute Director’s Fund, are gratefully acknowledged. Part of the research described in this paper was performed at the Canadian Light Source, a national research facility of the University of Saskatchewan, which is supported by the Canada Foundation for Innovation (CFI), the Natural Sciences and Engineering Research Council (NSERC), the National Research Council (NRC), the Canadian Institutes of Health Research (CIHR), the Government of Saskatchewan, and the University of Saskatchewan.

Dedication

Click to copy section linkSection link copied!

This work is dedicated to the memory of Professor Boris Žemva, whose passion for fluorine chemistry ignited our pursuit in this field.

References

Click to copy section linkSection link copied!

This article references 44 other publications.

  1. 1
    Broholm, C.; Cava, R. J.; Kivelson, S. A.; Nocera, D. G.; Norman, M. R.; Senthil, T. Quantum spin liquids. Science 2020, 367 (6475), eaay0668  DOI: 10.1126/science.aay0668
  2. 2
    Anderson, P. W. The Resonating Valence Bond State in La2CuO4 and Superconductivity. Science 1987, 235 (4793), 11961198,  DOI: 10.1126/science.235.4793.1196
  3. 3
    Grochala, W.; Hoffmann, R. Real and Hypothetical Intermediate-Valence AgII/AgIII and AgII/AgI Fluoride Systems as Potential Superconductors. Angew. Chem., Int. Ed. 2001, 40 (15), 27422781,  DOI: 10.1002/1521-3773(20010803)40:15<2742::AID-ANIE2742>3.0.CO;2-X
  4. 4
    Grochala, W. The theory-driven quest for a novel family of superconductors: fluorides. J. Mater. Chem. 2009, 19 (38), 69496968,  DOI: 10.1039/b904204k
  5. 5
    Žemva, B.; Hagiwara, R.; Casteel, W. J.; Lutar, K.; Jesih, A.; Bartlett, N. Spontaneous oxidation of xenon to Xe(II) by cationic Ag(II) in anhydrous hydrogen fluoride solutions. J. Am. Chem. Soc. 1990, 112 (12), 48464849,  DOI: 10.1021/ja00168a032
  6. 6
    Žemva, B. Protonic superacid anhydrous hydrogen fluoride as a solvent in the chemistry of high oxidation states. C. R. Acad. Sci., Ser. IIc: Chim. 1998, 1 (3), 151156,  DOI: 10.1016/S1387-1609(99)80073-5
  7. 7
    Grochala, W. Silverland: the Realm of Compounds of Divalent Silver─and Why They are Interesting. J. Supercond. Novel Magn. 2018, 31, 737752,  DOI: 10.1007/s10948-017-4326-8
  8. 8
    Gawraczyński, J.; Kurzydłowski, D.; Ewings, R. A.; Bandaru, S.; Gadomski, W.; Mazej, Z.; Ruani, G.; Bergenti, I.; Jaroń, T.; Ozarowski, A.; Hill, S.; Leszczyński, P. J.; Tokár, K.; Derzsi, M.; Barone, P.; Wohlfeld, K.; Lorenzana, J.; Grochala, W. Silver route to cuprate analogs. Proc. Natl. Acad. Sci. U.S.A. 2019, 116 (5), 14951500,  DOI: 10.1073/pnas.1812857116
  9. 9
    Grochala, W.; Egdell, R. G.; Edwards, P. P.; Mazej, Z.; Žemva, B. On the Covalency of Silver–Fluorine Bonds in Compounds of Silver(I), Silver(II) and Silver(III). ChemPhysChem 2003, 4 (9), 9971001,  DOI: 10.1002/cphc.200300777
  10. 10
    Bachar, N.; Koteras, K.; Gawraczynski, J.; Trzciński, W.; Paszula, J.; Piombo, R.; Barone, P.; Mazej, Z.; Ghiringhelli, G.; Nag, A.; Zhou, K.-J.; Lorenzana, J.; Van Der Marel, D.; Grochala, W. Charge-Transfer and dd excitations in AgF2. Phys. Rev. Res. 2022, 4 (2), 023108,  DOI: 10.1103/PhysRevResearch.4.023108
  11. 11
    Hettich, A. Über die Natur von Silbersubfluorid. Z. Anorg. Allg. Chem. 1927, 167 (1), 6774,  DOI: 10.1002/zaac.19271670106
  12. 12
    Terrey, H.; Diamond, H. The Crystal Structure of Silver Subfluoride. J. Chem. Soc. 1928, 28202824,  DOI: 10.1039/JR9280002820
  13. 13
    Andres, K.; Kuebler, N. A.; Robin, M. B. Superconductivity in Ag2F. J. Phys. Chem. Solids 1966, 27 (11–12), 17471748,  DOI: 10.1016/0022-3697(66)90104-1
  14. 14
    Ott, H. Die Strukturen von MnO, MnS, AgF, NiS, SnJ4, SrCl2, BaF2; Präzisionsmessungen einiger Alkalihalogenide. Z. Kristallogr. 1926, 63 (1–6), 222230,  DOI: 10.1524/zkri.1926.63.1.222
  15. 15
    Lozinšek, M.; Belak Vivod, M.; Dragomir, M. Crystal structure reinvestigation of silver(I) fluoride, AgF. IUCrData 2023, 8, x230018,  DOI: 10.1107/S2414314623000184
  16. 16
    Ebert, M. S.; Rodowskas, E. L.; Frazer, J. C. W. Higher valence states of silver. J. Am. Chem. Soc. 1933, 55 (7), 30563057,  DOI: 10.1021/ja01334a514
  17. 17
    Ruff, O.; Giese, M. Die Fluorierung des Silbers und Kupfers. Z. Anorg. Allg. Chem. 1934, 219 (2), 143148,  DOI: 10.1002/zaac.19342190206
  18. 18
    Jesih, A.; Lutar, K.; Žemva, B.; Bachmann, B.; Becker, S.; Müller, B. G.; Hoppe, R. Einkristalluntersuchungen an AgF2. Z. Anorg. Allg. Chem. 1990, 588 (1), 7783,  DOI: 10.1002/zaac.19905880110
  19. 19
    Bougon, R.; Bui Huy, T.; Lance, M.; Abazli, H. Synthesis and Properties of Silver Trifluoride, AgF3. Inorg. Chem. 1984, 23 (22), 36673668,  DOI: 10.1021/ic00190a049
  20. 20
    Žemva, B.; Lutar, K.; Jesih, A.; Casteel, W. J.; Wilkinson, A. P.; Cox, D. E.; von Dreele, R. B.; Borrmann, H.; Bartlett, N. Silver trifluoride: preparation, crystal structure, some properties, and comparison with AuF3. J. Am. Chem. Soc. 1991, 113 (11), 41924198,  DOI: 10.1021/ja00011a021
  21. 21
    Fischer, R.; Müller, B. G. Die Kristallstruktur von AgIIF[AgIIIF4]. Z. Anorg. Allg. Chem. 2002, 628 (12), 25922596,  DOI: 10.1002/1521-3749(200212)628:12<2592::AID-ZAAC2592>3.0.CO;2-O
  22. 22
    Graudejus, O.; Wilkinson, A. P.; Bartlett, N. Structural Features of Ag[AuF4] and Ag[AuF6] and the Structural Relationship of Ag[AgF4]2 and Au[AuF4]2 to Ag[AuF4]2. Inorg. Chem. 2000, 39 (7), 15451548,  DOI: 10.1021/ic991178t
  23. 23
    Grochala, W. On possible existence of pseudobinary mixed valence fluorides of Ag(I)/Ag(II): a DFT study. J. Mol. Model. 2011, 17, 22372248,  DOI: 10.1007/s00894-010-0949-4
  24. 24
    Kurzydłowski, D.; Derzsi, M.; Zurek, E.; Grochala, W. Fluorides of silver under large compression. Chem.─Eur. J. 2021, 27 (17), 55365545,  DOI: 10.1002/chem.202100028
  25. 25
    Rybin, N.; Chepkasov, I.; Novoselov, D. Y.; Anisimov, V. I.; Oganov, A. R. Prediction of Stable Silver Fluorides. J. Phys. Chem. C 2022, 126 (35), 1505715063,  DOI: 10.1021/acs.jpcc.2c04785
  26. 26
    Babel, D. Untersuchungen an ternären Fluoriden. III. Die Struktur des Na2CuF4. Z. Anorg. Allg. Chem. 1965, 336 (3–4), 200206,  DOI: 10.1002/zaac.19653360310
  27. 27
    Kurzydłowski, D.; Derzsi, M.; Budzianowski, A.; Jagličić, Z.; Koźmiński, W.; Mazej, Z.; Grochala, W. Polymorphism of Fluoroargentates(II): Facile Collapse of a Layered Network of α-K2AgF4 Due to the Insufficient Size of the Potassium Cation. Eur. J. Inorg. Chem. 2010, (19), 29192925,  DOI: 10.1002/ejic.201000124
  28. 28
    Baláž, P.; Achimovičová, M.; Baláž, M.; Billik, P.; Cherkezova-Zheleva, Z.; Criado, J. M.; Delogu, F.; Dutková, E.; Gaffet, E.; Gotor, F. J.; Kumar, R.; Mitov, I.; Rojac, T.; Senna, M.; Streletskii, A.; Wieczorek-Ciurowa, K. Hallmarks of mechanochemistry: from nanoparticles to technology. Chem. Soc. Rev. 2013, 42 (18), 75717637,  DOI: 10.1039/C3CS35468G
  29. 29
    Tan, D.; Garcia, F. Main group mechanochemistry: from curiosity to established protocols. Chem. Soc. Rev. 2019, 48 (8), 22742292,  DOI: 10.1039/C7CS00813A
  30. 30
    Friščić, T.; Mottillo, C.; Titi, H. M. Mechanochemistry for Synthesis. Angew. Chem., Int. Ed. 2020, 59 (3), 10181029,  DOI: 10.1002/anie.201906755
  31. 31
    Martinez, V.; Stolar, T.; Karadeniz, B.; Brekalo, I.; Užarević, K. Advancing mechanochemical synthesis by combining milling with different energy sources. Nat. Rev. Chem 2023, 7, 5165,  DOI: 10.1038/s41570-022-00442-1
  32. 32
    Šepelák, V.; Düvel, A.; Wilkening, M.; Becker, K. D.; Heitjans, P. Mechanochemical reactions and syntheses of oxides. Chem. Soc. Rev. 2013, 42 (18), 75077520,  DOI: 10.1039/c2cs35462d
  33. 33
    Preishuber-Pflügl, F.; Wilkening, M. Mechanochemically synthesized fluorides: local structures and ion transport. Dalton Trans. 2016, 45 (21), 86758687,  DOI: 10.1039/C6DT00944A
  34. 34
    Ruprecht, B.; Wilkening, M.; Feldhoff, A.; Steuernagel, S.; Heitjans, P. High anion conductivity in a ternary non-equilibrium phase of BaF2 and CaF2 with mixed cations. Phys. Chem. Chem. Phys. 2009, 11 (17), 30713081,  DOI: 10.1039/b901293a
  35. 35
    Kurzydłowski, D.; Jaroń, T.; Ozarowski, A.; Hill, S.; Jagličić, Z.; Filinchuk, Y.; Mazej, Z.; Grochala, W. Local and Cooperative Jahn–Teller Effect and Resultant Magnetic Properties of M2AgF4 (M = Na–Cs) Phases. Inorg. Chem. 2016, 55 (21), 1147911489,  DOI: 10.1021/acs.inorgchem.6b02037
  36. 36
    Lozinšek, M.; Goreshnik, G.; Žemva, B. Silver(I) Tetrafluoridooxidovanadate(V) – Ag[VOF4]. Acta Chim. Slov. 2014, 61 (3), 542547
  37. 37
    Brown, I. D. Recent Developments in the Methods and Applications of the Bond Valence Model. Chem. Rev. 2009, 109 (12), 68586919,  DOI: 10.1021/cr900053k
  38. 38
    Gawraczyński, J. Optical Spectroscopy of Selected Divalent Silver Compounds. Ph.D. Dissertation, University of Warsaw, 2019.
  39. 39
    Bonner, J. C.; Fisher, M. E. Linear Magnetic Chains with Anisotropic Coupling. Phys. Rev. 1964, 135 (3A), A640A658,  DOI: 10.1103/PhysRev.135.A640
  40. 40
    Tong, J.; Lee, C.; Whangbo, M. H.; Kremer, R. K.; Simon, A.; Köhler, J. Cooperative Jahn–Teller distortion leading to the spin-1/2 uniform antiferromagnetic chains in triclinic perovskites AgCuF3 and NaCuF3. Solid State Sci. 2010, 12 (5), 680684,  DOI: 10.1016/j.solidstatesciences.2009.02.028
  41. 41
    Odenthal, R. H.; Hoppe, R. Fluoroargentate(II) der Alkalimetalle. Monatsh. Chem. 1971, 102, 13401350,  DOI: 10.1007/BF00917190
  42. 42
    Kaiser, V.; Otto, M.; Binder, F.; Babel, D. Jahn-Teller-Effekt und Kristallstruktur-Verzerrung bei den Kupfer-Fluorperowskiten NaCuF3 und RbCuF3. Z. Anorg. Allg. Chem. 1990, 585 (1), 93104,  DOI: 10.1002/zaac.19905850112
  43. 43
    Mazej, Z.; Goreshnik, E.; Jagličić, Z.; Gaweł, B.; Łasocha, W.; Grzybowska, D.; Jaroń, T.; Kurzydłowski, D.; Malinowski, P.; Koźminski, W.; Szydłowska, J.; Leszczyński, P.; Grochala, W. KAgF3, K2AgF4 and K3Ag2F7: important steps towards a layered antiferromagnetic fluoroargentate(II). CrystEngComm 2009, 11 (8), 17021710,  DOI: 10.1039/B902161B
  44. 44
    Kurzydłowski, D.; Mazej, Z.; Jagličić, Z.; Filinchuk, Y.; Grochala, W. Structural transition and unusually strong antiferromagnetic superexchange coupling in perovskite KAgF3. Chem. Commun. 2013, 49 (56), 62626264,  DOI: 10.1039/c3cc41521j

Cited By

Click to copy section linkSection link copied!

This article has not yet been cited by other publications.

Journal of the American Chemical Society

Cite this: J. Am. Chem. Soc. 2024, 146, 44, 30510–30517
Click to copy citationCitation copied!
https://doi.org/10.1021/jacs.4c11772
Published October 24, 2024

Copyright © 2024 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY 4.0 .

Article Views

1495

Altmetric

-

Citations

-
Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Figure 1

    Figure 1. (a) Rietveld refinement of the laboratory PXRD data (Ag Kα radiation, λ = 0.5609 Å) of mechanochemically synthesized sample of AgI2AgIIF4 collected at room temperature. Quantitative analysis: AgI2AgIIF4: 88.7(2) wt %; AgF: 11.3(6) wt %. (b) Three-phase Rietveld refinement of the synchrotron data collected on the AgI2AgIIF4 sample prepared by solid-state method. Quantitative analysis: AgI2AgIIF4: 76.2(1) wt %; AgF: 22.9(1) wt %; AgF2: 0.9(5) wt %. (c) Unit cell of the AgI2AgIIF4 (Ag3F4) crystal structure and a chain of edge-sharing [AgIIF2F4/2]2– distorted octahedral units with 4 + 2 coordination, which extends parallel to a-crystallographic axis. (d) The 5 K NPD pattern of the ball-milled AgI2AgIIF4 sample showing that the only impurity is small amounts of diamagnetic AgF. Quantitative analysis: AgI2AgIIF4: 89.4(1) wt %; AgF: 10.6(1) wt %.

    Figure 2

    Figure 2. (a) Raman spectra of AgI2AgIIF4 and β-K2AgF4 synthesized by ball milling. (b) Temperature-dependent magnetic susceptibility of mechanochemically prepared AgI2AgIIF4 sample measured in a magnetic field of 1 kOe. The inset shows the Curie–Weiss fit to the inverse of magnetic susceptibility in the 15–300 K range. (c) Zero field cooled (ZFC) and field cooled (FC) data in the 2–25 K temperature region showing the upturn in the magnetic susceptibility characteristic of AFM, while the Bonner–Fisher fit for the 2–15 K region is shown in the inset. (d) Magnetization of AgI2AgIIF4 as a function of magnetic field measured at 2 and 5 K. The hysteresis shown in the inset indicates a FM ground state.

    Figure 3

    Figure 3. (a) Rietveld refinement of the laboratory PXRD data (Ag Kα radiation, λ = 0.5609 Å) of mechanochemically synthesized sample of AgIAgIIF3, measured at room temperature. Quantitative analysis: AgIAgIIF3: 95.6(1) wt %; AgI2AgIIF4: 3.3(2) wt %; AgF2: 1.1(3) wt %. (b) Unit cell of the AgIAgIIF3 (Ag2F3) crystal structure with apex-sharing [AgIIF6/2] distorted octahedral units with 4 + 2 coordination. (c) Raman spectrum of ball-milled AgIAgIIF3. For comparison, the Raman spectrum of KAgF3 is also shown. (d) Weight fraction of phases in the ball-milled samples obtained from Rietveld refinement of the laboratory PXRD data as a function of milling time. The amorphous phase was not accounted in the refinement; the errors are smaller than the symbols in the graph.

    Figure 4

    Figure 4. Temperature-dependent magnetic susceptibility, χ, of AgIAgIIF3, measured in a magnetic field of (a) 1 and (b) 10 kOe. The inset in (a) shows the first derivative indicating a deviation from the Curie–Weiss law at 7.3(7) K. The inset in (b) shows the inverse of the magnetic susceptibility (FC) measured at 10 kOe. (c) Plot of the effective magnetic moment, μeff, as a function of temperature for FC and ZFC curves at 1 kOe and for the FC curve measured at 10 kOe. (d) Magnetization of AgIAgIIF3 as a function of magnetic field, measured at 2 K.

  • References


    This article references 44 other publications.

    1. 1
      Broholm, C.; Cava, R. J.; Kivelson, S. A.; Nocera, D. G.; Norman, M. R.; Senthil, T. Quantum spin liquids. Science 2020, 367 (6475), eaay0668  DOI: 10.1126/science.aay0668
    2. 2
      Anderson, P. W. The Resonating Valence Bond State in La2CuO4 and Superconductivity. Science 1987, 235 (4793), 11961198,  DOI: 10.1126/science.235.4793.1196
    3. 3
      Grochala, W.; Hoffmann, R. Real and Hypothetical Intermediate-Valence AgII/AgIII and AgII/AgI Fluoride Systems as Potential Superconductors. Angew. Chem., Int. Ed. 2001, 40 (15), 27422781,  DOI: 10.1002/1521-3773(20010803)40:15<2742::AID-ANIE2742>3.0.CO;2-X
    4. 4
      Grochala, W. The theory-driven quest for a novel family of superconductors: fluorides. J. Mater. Chem. 2009, 19 (38), 69496968,  DOI: 10.1039/b904204k
    5. 5
      Žemva, B.; Hagiwara, R.; Casteel, W. J.; Lutar, K.; Jesih, A.; Bartlett, N. Spontaneous oxidation of xenon to Xe(II) by cationic Ag(II) in anhydrous hydrogen fluoride solutions. J. Am. Chem. Soc. 1990, 112 (12), 48464849,  DOI: 10.1021/ja00168a032
    6. 6
      Žemva, B. Protonic superacid anhydrous hydrogen fluoride as a solvent in the chemistry of high oxidation states. C. R. Acad. Sci., Ser. IIc: Chim. 1998, 1 (3), 151156,  DOI: 10.1016/S1387-1609(99)80073-5
    7. 7
      Grochala, W. Silverland: the Realm of Compounds of Divalent Silver─and Why They are Interesting. J. Supercond. Novel Magn. 2018, 31, 737752,  DOI: 10.1007/s10948-017-4326-8
    8. 8
      Gawraczyński, J.; Kurzydłowski, D.; Ewings, R. A.; Bandaru, S.; Gadomski, W.; Mazej, Z.; Ruani, G.; Bergenti, I.; Jaroń, T.; Ozarowski, A.; Hill, S.; Leszczyński, P. J.; Tokár, K.; Derzsi, M.; Barone, P.; Wohlfeld, K.; Lorenzana, J.; Grochala, W. Silver route to cuprate analogs. Proc. Natl. Acad. Sci. U.S.A. 2019, 116 (5), 14951500,  DOI: 10.1073/pnas.1812857116
    9. 9
      Grochala, W.; Egdell, R. G.; Edwards, P. P.; Mazej, Z.; Žemva, B. On the Covalency of Silver–Fluorine Bonds in Compounds of Silver(I), Silver(II) and Silver(III). ChemPhysChem 2003, 4 (9), 9971001,  DOI: 10.1002/cphc.200300777
    10. 10
      Bachar, N.; Koteras, K.; Gawraczynski, J.; Trzciński, W.; Paszula, J.; Piombo, R.; Barone, P.; Mazej, Z.; Ghiringhelli, G.; Nag, A.; Zhou, K.-J.; Lorenzana, J.; Van Der Marel, D.; Grochala, W. Charge-Transfer and dd excitations in AgF2. Phys. Rev. Res. 2022, 4 (2), 023108,  DOI: 10.1103/PhysRevResearch.4.023108
    11. 11
      Hettich, A. Über die Natur von Silbersubfluorid. Z. Anorg. Allg. Chem. 1927, 167 (1), 6774,  DOI: 10.1002/zaac.19271670106
    12. 12
      Terrey, H.; Diamond, H. The Crystal Structure of Silver Subfluoride. J. Chem. Soc. 1928, 28202824,  DOI: 10.1039/JR9280002820
    13. 13
      Andres, K.; Kuebler, N. A.; Robin, M. B. Superconductivity in Ag2F. J. Phys. Chem. Solids 1966, 27 (11–12), 17471748,  DOI: 10.1016/0022-3697(66)90104-1
    14. 14
      Ott, H. Die Strukturen von MnO, MnS, AgF, NiS, SnJ4, SrCl2, BaF2; Präzisionsmessungen einiger Alkalihalogenide. Z. Kristallogr. 1926, 63 (1–6), 222230,  DOI: 10.1524/zkri.1926.63.1.222
    15. 15
      Lozinšek, M.; Belak Vivod, M.; Dragomir, M. Crystal structure reinvestigation of silver(I) fluoride, AgF. IUCrData 2023, 8, x230018,  DOI: 10.1107/S2414314623000184
    16. 16
      Ebert, M. S.; Rodowskas, E. L.; Frazer, J. C. W. Higher valence states of silver. J. Am. Chem. Soc. 1933, 55 (7), 30563057,  DOI: 10.1021/ja01334a514
    17. 17
      Ruff, O.; Giese, M. Die Fluorierung des Silbers und Kupfers. Z. Anorg. Allg. Chem. 1934, 219 (2), 143148,  DOI: 10.1002/zaac.19342190206
    18. 18
      Jesih, A.; Lutar, K.; Žemva, B.; Bachmann, B.; Becker, S.; Müller, B. G.; Hoppe, R. Einkristalluntersuchungen an AgF2. Z. Anorg. Allg. Chem. 1990, 588 (1), 7783,  DOI: 10.1002/zaac.19905880110
    19. 19
      Bougon, R.; Bui Huy, T.; Lance, M.; Abazli, H. Synthesis and Properties of Silver Trifluoride, AgF3. Inorg. Chem. 1984, 23 (22), 36673668,  DOI: 10.1021/ic00190a049
    20. 20
      Žemva, B.; Lutar, K.; Jesih, A.; Casteel, W. J.; Wilkinson, A. P.; Cox, D. E.; von Dreele, R. B.; Borrmann, H.; Bartlett, N. Silver trifluoride: preparation, crystal structure, some properties, and comparison with AuF3. J. Am. Chem. Soc. 1991, 113 (11), 41924198,  DOI: 10.1021/ja00011a021
    21. 21
      Fischer, R.; Müller, B. G. Die Kristallstruktur von AgIIF[AgIIIF4]. Z. Anorg. Allg. Chem. 2002, 628 (12), 25922596,  DOI: 10.1002/1521-3749(200212)628:12<2592::AID-ZAAC2592>3.0.CO;2-O
    22. 22
      Graudejus, O.; Wilkinson, A. P.; Bartlett, N. Structural Features of Ag[AuF4] and Ag[AuF6] and the Structural Relationship of Ag[AgF4]2 and Au[AuF4]2 to Ag[AuF4]2. Inorg. Chem. 2000, 39 (7), 15451548,  DOI: 10.1021/ic991178t
    23. 23
      Grochala, W. On possible existence of pseudobinary mixed valence fluorides of Ag(I)/Ag(II): a DFT study. J. Mol. Model. 2011, 17, 22372248,  DOI: 10.1007/s00894-010-0949-4
    24. 24
      Kurzydłowski, D.; Derzsi, M.; Zurek, E.; Grochala, W. Fluorides of silver under large compression. Chem.─Eur. J. 2021, 27 (17), 55365545,  DOI: 10.1002/chem.202100028
    25. 25
      Rybin, N.; Chepkasov, I.; Novoselov, D. Y.; Anisimov, V. I.; Oganov, A. R. Prediction of Stable Silver Fluorides. J. Phys. Chem. C 2022, 126 (35), 1505715063,  DOI: 10.1021/acs.jpcc.2c04785
    26. 26
      Babel, D. Untersuchungen an ternären Fluoriden. III. Die Struktur des Na2CuF4. Z. Anorg. Allg. Chem. 1965, 336 (3–4), 200206,  DOI: 10.1002/zaac.19653360310
    27. 27
      Kurzydłowski, D.; Derzsi, M.; Budzianowski, A.; Jagličić, Z.; Koźmiński, W.; Mazej, Z.; Grochala, W. Polymorphism of Fluoroargentates(II): Facile Collapse of a Layered Network of α-K2AgF4 Due to the Insufficient Size of the Potassium Cation. Eur. J. Inorg. Chem. 2010, (19), 29192925,  DOI: 10.1002/ejic.201000124
    28. 28
      Baláž, P.; Achimovičová, M.; Baláž, M.; Billik, P.; Cherkezova-Zheleva, Z.; Criado, J. M.; Delogu, F.; Dutková, E.; Gaffet, E.; Gotor, F. J.; Kumar, R.; Mitov, I.; Rojac, T.; Senna, M.; Streletskii, A.; Wieczorek-Ciurowa, K. Hallmarks of mechanochemistry: from nanoparticles to technology. Chem. Soc. Rev. 2013, 42 (18), 75717637,  DOI: 10.1039/C3CS35468G
    29. 29
      Tan, D.; Garcia, F. Main group mechanochemistry: from curiosity to established protocols. Chem. Soc. Rev. 2019, 48 (8), 22742292,  DOI: 10.1039/C7CS00813A
    30. 30
      Friščić, T.; Mottillo, C.; Titi, H. M. Mechanochemistry for Synthesis. Angew. Chem., Int. Ed. 2020, 59 (3), 10181029,  DOI: 10.1002/anie.201906755
    31. 31
      Martinez, V.; Stolar, T.; Karadeniz, B.; Brekalo, I.; Užarević, K. Advancing mechanochemical synthesis by combining milling with different energy sources. Nat. Rev. Chem 2023, 7, 5165,  DOI: 10.1038/s41570-022-00442-1
    32. 32
      Šepelák, V.; Düvel, A.; Wilkening, M.; Becker, K. D.; Heitjans, P. Mechanochemical reactions and syntheses of oxides. Chem. Soc. Rev. 2013, 42 (18), 75077520,  DOI: 10.1039/c2cs35462d
    33. 33
      Preishuber-Pflügl, F.; Wilkening, M. Mechanochemically synthesized fluorides: local structures and ion transport. Dalton Trans. 2016, 45 (21), 86758687,  DOI: 10.1039/C6DT00944A
    34. 34
      Ruprecht, B.; Wilkening, M.; Feldhoff, A.; Steuernagel, S.; Heitjans, P. High anion conductivity in a ternary non-equilibrium phase of BaF2 and CaF2 with mixed cations. Phys. Chem. Chem. Phys. 2009, 11 (17), 30713081,  DOI: 10.1039/b901293a
    35. 35
      Kurzydłowski, D.; Jaroń, T.; Ozarowski, A.; Hill, S.; Jagličić, Z.; Filinchuk, Y.; Mazej, Z.; Grochala, W. Local and Cooperative Jahn–Teller Effect and Resultant Magnetic Properties of M2AgF4 (M = Na–Cs) Phases. Inorg. Chem. 2016, 55 (21), 1147911489,  DOI: 10.1021/acs.inorgchem.6b02037
    36. 36
      Lozinšek, M.; Goreshnik, G.; Žemva, B. Silver(I) Tetrafluoridooxidovanadate(V) – Ag[VOF4]. Acta Chim. Slov. 2014, 61 (3), 542547
    37. 37
      Brown, I. D. Recent Developments in the Methods and Applications of the Bond Valence Model. Chem. Rev. 2009, 109 (12), 68586919,  DOI: 10.1021/cr900053k
    38. 38
      Gawraczyński, J. Optical Spectroscopy of Selected Divalent Silver Compounds. Ph.D. Dissertation, University of Warsaw, 2019.
    39. 39
      Bonner, J. C.; Fisher, M. E. Linear Magnetic Chains with Anisotropic Coupling. Phys. Rev. 1964, 135 (3A), A640A658,  DOI: 10.1103/PhysRev.135.A640
    40. 40
      Tong, J.; Lee, C.; Whangbo, M. H.; Kremer, R. K.; Simon, A.; Köhler, J. Cooperative Jahn–Teller distortion leading to the spin-1/2 uniform antiferromagnetic chains in triclinic perovskites AgCuF3 and NaCuF3. Solid State Sci. 2010, 12 (5), 680684,  DOI: 10.1016/j.solidstatesciences.2009.02.028
    41. 41
      Odenthal, R. H.; Hoppe, R. Fluoroargentate(II) der Alkalimetalle. Monatsh. Chem. 1971, 102, 13401350,  DOI: 10.1007/BF00917190
    42. 42
      Kaiser, V.; Otto, M.; Binder, F.; Babel, D. Jahn-Teller-Effekt und Kristallstruktur-Verzerrung bei den Kupfer-Fluorperowskiten NaCuF3 und RbCuF3. Z. Anorg. Allg. Chem. 1990, 585 (1), 93104,  DOI: 10.1002/zaac.19905850112
    43. 43
      Mazej, Z.; Goreshnik, E.; Jagličić, Z.; Gaweł, B.; Łasocha, W.; Grzybowska, D.; Jaroń, T.; Kurzydłowski, D.; Malinowski, P.; Koźminski, W.; Szydłowska, J.; Leszczyński, P.; Grochala, W. KAgF3, K2AgF4 and K3Ag2F7: important steps towards a layered antiferromagnetic fluoroargentate(II). CrystEngComm 2009, 11 (8), 17021710,  DOI: 10.1039/B902161B
    44. 44
      Kurzydłowski, D.; Mazej, Z.; Jagličić, Z.; Filinchuk, Y.; Grochala, W. Structural transition and unusually strong antiferromagnetic superexchange coupling in perovskite KAgF3. Chem. Commun. 2013, 49 (56), 62626264,  DOI: 10.1039/c3cc41521j
  • Supporting Information

    Supporting Information


    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/jacs.4c11772.

    • The experimental procedures, description of the characterization techniques, and results of analyses (PDF)

    Accession Codes

    Deposition numbers 2321112 (for Ag2F3) and 2321184 (for Ag3F4) contain the supplementary crystallographic data for this paper. These data are provided free of charge by the joint Cambridge Crystallographic Data Centre and Fachinformationszentrum Karlsruhe Access Structures service.


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.