Open Access Article
Lina
Cherni
a,
Karin
El Rifaii
b,
Henricus H.
Wensink
*b,
Sarah M.
Chevrier
a,
Claire
Goldmann
b,
Laurent J.
Michot
c,
Patrick
Davidson
*b and
Jean-Christophe P.
Gabriel
*a
aUniversité Paris-Saclay, CEA, CNRS, NIMBE-LICSEN, 91191 Gif-sur-Yvette, France. E-mail: Jean-Christophe.gabriel@cea.fr
bLaboratoire de Physique des Solides, Université Paris-Saclay, CNRS, 91405 Orsay, France. E-mail: rik.wensink@universite-paris-saclay.fr; patrick.davidson@universite-paris-saclay.fr
cLaboratory of Physical Chemistry of Electrolytes and Interfacial Nanosystems (PHENIX), Sorbonne Université, CNRS, 75005 Paris, France
First published on 6th November 2023
We report here the highly ordered restacking of the layered phosphatoantimonic dielectric materials H3(1−x)M3xSb3P2O14, (where M = Li, Na, K, Rb, Cs and 0 ≤ x ≤ 1), from their nanosheets dispersed in colloidal suspension, induced by a simple pH change using alkaline bases. H3Sb3P2O14 aqueous suspensions are some of the rare examples of colloidal suspensions based on 2D materials exhibiting a lamellar liquid crystalline phase. Because the lamellar period can reach several hundred nanometers, the suspensions show vivid structural colors and because these colors are sensitive to various chemicals, the suspensions can be used as sensors. The structures of the lamellar liquid crystalline phase and the restacked phase have been studied by X-ray scattering (small and wide angle), which has followed the dependence of the lamellar/restacked phase equilibrium on the cation exchange rate, x. The X-ray diffraction pattern of the restacked phase is almost identical to that of the M3Sb3P2O14 crystalline phase, showing that the restacking is highly accurate and avoids the turbostratic disorder of the nanosheets classically observed in nanosheet stacking of other 2D materials. Strikingly, the restacking process exhibits features highly reminiscent of a first-order phase transition, with the existence of a phase coexistence region where both ∼1 nm (interlayer spacing of the restacked phase) and ∼120 nm lamellar periods can be observed simultaneously. Furthermore, this first-order phase transition is well described theoretically by incorporating a Lennard-Jones-type lamellar interaction potential into an entropy-based statistical physics model of the lamellar phase of nanosheets. Our work shows that the precise cation exchange produced at room temperature by a classical neutralization reaction using alkaline bases leads to a crystal-like restacking of the exfoliated free Sb3P2O143− nanosheets from suspension, avoiding the turbostratic disorder typical of van der Waals 2D materials, which is detrimental to the controlled deposition of nanosheets into complex integrated electronic, spintronic, photonic or quantum structures.
We therefore report here our investigations into this issue by studying the restacking of the dielectric 2D material M3Sb3P2O14 (hereafter called M3) (M = H, Li, Na, K, Rb, Cs), a well-defined model system of low-dimensional compounds.12,13 Furthermore, this material has been shown to readily exfoliate in water, resulting in colloidal suspensions of individual nanosheets that exhibit a lamellar liquid crystalline phase at very low mass fractions (≈1 w/w%) and a rare example of a lamellar material that diffracts visible light.14,15 As a result, aqueous suspensions of H3Sb3P2O14 (H3) have structural colors that vary from blue to red depending on concentration and chemical or physical stress,16 which also makes this material a good candidate for photonic and sensing applications.17,18 Therefore, we report the observation, in natural and polarized light, of the aspect of lamellar liquid crystalline phases in different systems when varying the cation exchange rate (x) and the alkaline metal M. We then present our X-ray scattering study of the structural evolution of the nanosheet suspensions and their restacking. Finally, we discuss and theoretically model the transition from the exfoliated to the restacked state from a thermodynamic point of view.
| H3Sb3P2O14 + 3xMOH → H3(1−x)M3xSb3P2O14 + 3xH2O | (1) |
![]() | ||
| Fig. 1 Potentiometric titration curves of H3Sb3P2O14 dispersions with different alkaline base solutions: LiOH, NaOH, KOH, RbOH, and CsOH. x (top axis) is the theoretical exchange rate. | ||
In addition, the observation of all samples between crossed polarizers (Fig. 2) revealed that the birefringence typical of the lamellar phase gradually disappears with increasing exchange rate for the different H3(1−x)M3x systems. However, a small amount of lamellar phase may persist up to high values of x as a thin birefringent layer localized between the precipitate and the upper isotropic phase.
For the H3(1−x)Li3x system, at x = 1, a single white precipitate phase is found at the bottom of the tube, coexisting with a very dilute isotropic suspension of nanosheets at the top. This indicates that the exchange of H+ cations by Li+ is complete and that the phase at the bottom of the tube corresponds to the water-insoluble solid Li3Sb3P2O14. The precipitate begins to appear at x ≈ 0.6, as the liquid-crystalline lamellar phase begins to disappear.
For the other H3(1−x)M3x systems, with M = Na, K, and Rb, the values at which the precipitate appeared were lower, at x ≈ 0.2, but this was associated with the observation of a wider range of x values in which both the lamellar phase and the precipitate coexisted. Finally, for the H3(1−x)Cs3x system, the precipitation did not start until x = 0.5 (the maximum x value studied in this study), at which point the lamellar phase, with its blue characteristic color, could still be observed.
![]() | ||
| Fig. 3 SAXS (left) and WAXS (right) patterns of the bottom phase of the H3(1−x)K3x samples shown in Fig. 2. The peaks related to the K3Sb3P2O14 crystalline phase are shown with red dashed lines. The peak tagged with a star (*) is the signature of the lamellar phase in WAXS patterns. | ||
In the WAXS patterns (Fig. 3, right), very sharp diffraction peaks start to appear at x = 0.2 and become more and more intense upon increasing x values. All the peaks present in the WAXS patterns have been identified as they actually correspond to the diffraction peaks of the crystalline K3Sb3P2O14. 1.32H2O powder observed long ago during the determination of its crystallographic structure.12,20 Therefore, quite surprisingly, the white precipitate observed in Fig. 2 at high values of x is simply made of K3Sb3P2O14 hydrated crystallites and the flocculation of the colloidal suspension leads to a highly organized state of the nanosheets.
The WAXS patterns also display an additional peak at q = 10 nm−1 that is present for all values of x. This peak, which is strongest for x = 0 (i.e. in the pure liquid-crystalline lamellar phase), represents the first (and strongest) reflection arising from the 2-dimensional crystalline order of each individual nanosheet. It is therefore the signature in WAXS of the lamellar mesophase. Its intensity decreases with increasing x that is with increasing proportion of the crystalline precipitate. Altogether, the X-ray scattering data therefore confirm the visual observations described above.
The lamellar period, d, of the mesophase is related to the position of the first-order diffraction peak in SAXS by Bragg's law, d = 2π/q. Similarly, the stacking period of the nanosheets in the crystalline precipitate is also directly obtained by the position of the (003) reflection in WAXS since there are three nanosheets per crystal cell along the c-axis. The dependence of these periods on the cation exchange rate, x, for both phases of the H3Sb3P2O14 system, is presented in Fig. 4. For x ≤ 0.4, the lamellar period increases from 119 nm at x = 0 to 165 nm at x = 0.4, which is consistent with the color of the samples that slowly shifts from blue at x = 0 to green at x = 0.4. In contrast, the stacking period of the restacked crystalline phase remains remarkably constant at 1 nm. Between x = 0.2 and x = 0.45, the lamellar liquid-crystalline phase and the restacked crystalline phase coexist. At x = 0.5, the remaining amount of lamellar phase is negligible.
Very importantly, none of the many samples investigated in this study showed a lamellar period intermediate between 1 and 100 nm, which suggests an abrupt transition between the lamellar mesophase and the restacked crystalline state, with phase coexistence, i.e. a 1st order phase transition that calls for theoretical modeling.
One may wonder why the restacked phases show such good crystallinity instead of turbostratic disorder. This can be explained by the corrugated surface of the Sb3P2O143− nanosheets, which is in stark contrast to flat 2D nanosheets, such as graphene. Indeed, the crystallographic structure of the former, presented in Fig. 5, shows that: (i) PO4 tetrahedra are alternately located above or below a plane of SbO6 octahedra; (ii) these tetrahedra are themselves centered in small cavities of the SbO6 plane which are located at the centers of the 6-ring structure of SbO6 octahedra. Such a structure creates a lock-in mechanism that allows only for a single stacking orientation of the nanosheets (modulo 2π/3). This is further detailed with a quick animation in ESI II.† Such mechanism cannot take place for common van der Waals 2D materials, which usually do not present any surface corrugation or structural roughness to play this role. To further investigate the generality of this topologically induced restacking mechanism and associated crystallinity, we searched the literature for other such examples. It is worth noting that in the only similar report of crystallinity induced by re-stacking, although the mechanism was not discussed, if one looks at the structure of the nanosheets involved, one can find similar protruding tetrahedra.21
FL ∼ (M − 1)AfL exp(−d/λD) | (2) |
Fintra ∼ kBTM(ln ρ⊥Λ2 − 1) | (3) |
![]() | (4) |
Minimizing with respect to d and implicitly renormalizing all variables in terms of the unit length λD and energy kBT, we obtain the equilibrium lamellar distance as a solution of the non-linear equation:
![]() | (5) |
![]() | (6) |
the intralamellar packing fraction of discs with diameter D which from now on serves as our unit of length. Furthermore, a = π
BλDwL is a prefactor in terms of the Bjerrum length
B and electrostatic amplitude wL that gathers various characteristics pertaining to the effective charge, shape, and intralamellar fluid structure of the nanosheets.16 Furthermore, for the interlamellar repulsion we identify fL = 8/π
BλD (in dimensionless units). Taking typical values for the Sb3P2O143− nanosheets with average diameter D = 〈D〉 ≈ 800 nm we find λD ≈ 0.125,
B ≈ 9 × 10−4, a ≈ 1.7 × 104 and fL ≈ 2.3 × 104.16
In the restacked state, the long-range electrostatic forces are no longer dominant and we need to account for additional (van der Waals) attractive and steep repulsive steric forces that are at play at short interlamellar separations. We thus write:
![]() | (7) |
For the intralamellar part, we combine the ideal gas contribution eqn (2) with the excess part eqn (5) that accounts for disc-disc correlations due to electrostatic repulsion and hard-core volume exclusion:16
![]() | (8) |
The free energy diverges when the discs reach close-packing, ϕ⊥ → 1, even though actual close packing of uniform-sized discs happens at
. We assume that the van der Waals attraction does not affect the intralamellar structure given the vanishing side-by-side contact area between two co-planar sheets. As before, minimization of the total free energy Ftot = FL + Fintra with respect to ∂Ftot/∂d = 0 yields the equilibrium lamellar distance which now further depends on the lamellar Lennard-Jones amplitude ε and range d* (see ESI†).
The possibility of an isostructural lamellar phase co-existence induced by the van der Waals forces can be explored by plotting (van der Waals) isotherms in terms of the total osmotic pressure of the lamellar phase P/ρL2 = ∂(Ftot/N)/∂ρL upon variation of the LLJ amplitude ε that may be interpreted as an inverse effective temperature. The isotherms in Fig. 6 clearly demonstrate the possibility of a gas–liquid type coexistence between a high-density restacked lamellar phase and a swollen lamellar phase (indicated by the dotted lines). For the particular case analysed in Fig. 4, we find that the lamellar system remains uniform for weak van der Waals forces (ε < 0.1) but phase separates at elevated attraction ε ≫ 1.
From a physical chemistry point of view, this study sheds new light on the exfoliation/restacking process of nanosheets in colloidal suspension, which should be considered as a first-order equilibrium phase transition. From the more applied perspective of controlled deposition of 2D materials into complex integrated structures, our investigation shows that the use of self-locking 2D materials provides a new means to avoid the undesired turbostratic orientation disorder of the nanosheets.
K3Sb3P2O14 was first synthesized using a stoichiometric mixture of NH4H2PO4 (0.224 mol, 25.823 g), Sb2O3 (MERCK, 0.166 mol, 48.367 g) and KNO3 (PROLABO, 0.332 mol, 33.552 g) that was heated at 300 °C for 10 h, then at 1000 °C for 24 h, using a ramp of 50 °C h−1. X-ray powder diffraction showed that the resulting solid (K3Sb3P2O14) had an impurity level lower than 1%.12
A cationic exchange of K3Sb3P2O14 by acidic treatment was done to produce the phosphatoantimonic acid H3Sb3P2O14. For this, 60 g of K3Sb3P2O14 was dissolved and stirred in a 2 L solution of 7.5 M nitric acid at 50 °C for 24 h. Centrifugation at 4500g was made to recover the treated solid. This procedure was repeated three times to complete the cationic exchange.
The solid obtained (H3Sb3P2O14) was finally rinsed several times with absolute ethanol and recovered by centrifugation, then dried in air at 60 °C overnight.
A solution of H3Sb3P2O14 (solution A) in water was first prepared by dispersing and stirring 20 g of the H3Sb3P2O14 powder into 1 L of 18.2 MΩ cm H2O (PURELAB® Chorus, ELGA, 18.2 MΩ cm) for 30 min. The solution A was then centrifuged at 3600 G for 15 min and the supernatant, which appeared milky and opalescent, was collected (solution B), whose concentration, measured by thermogravimetric analysis (TGA), is 1.65 w/w%. Solution B's residual nitrate ion concentration, estimated with a JBL nitrate test, was below 1 ppm (lowest level of detection of the test). Another solution of H3Sb3P2O14 was prepared by diluting solution B by a factor of 10 (solution C, at 0.165 w/w%).
First, for each targeted value of x, the volume of the base required was added drop by drop to 5 mL of H3Sb3P2O14 suspension (solution B) under vigorous stirring to obtain a homogeneous mixture. Then, sample series with different dilution factors (1, 1.25, 2, 3, 5, and 10) were prepared from each of these mixtures by adding the required volume of water (18.2 MΩ cm). All samples are therefore defined by the nanosheet weight fraction (in w/w%) and the exchange rate, x. Each sample prepared had a final volume of 1 mL and was poured in a 2 mL glass vial.
In a typical titration of H3Sb3P2O14, 15 mL of H3Sb3P2O14 suspension (solution C, 0.165 w/w%) was titrated by adding the selected basic solution at a low rate of 0.05 mL min−1, by increment of 0.01 mL, under vigorous stirring. Titration experiments were performed at very slow rates of base addition because the titration of an acidic colloidal suspension of charged nanosheets is a complex process. Indeed, each time a drop of basic solution is added into the suspension, before the base neutralizes the acid, a charge excess builds up locally and may screen the electrostatic repulsions that provide the nanosheets with colloidal stability, and therefore may lead to flocculation around the drop. Prior and after each titration, the pH electrode was rinsed with deionized water.
sin
θ)/λ where 2θ is the scattering angle) ranged between 10−2 and 0.5 nm−1. The beam size was 375 × 75 μm2 at the sample level. The scattering patterns were recorded with an Eiger-4 M detector and the exposure time was typically 0.5 s. The usual SAXS data reduction procedures were made, using the Swing data reduction software (Foxtrot 3.4.9), and the data was displayed either as 2-dimensional SAXS patterns or as plots of the scattered intensity versus scattering vector modulus, I(q), obtained by azimuthal averaging of the SAXS patterns. The samples were filled into cylindrical Lindemann glass capillaries of 1.0 ± 0.1 mm diameter (Glas-Technik & Konstruktion, Germany) and sealed with silicone glue.
To facilitate the indexation of the peaks of the WAXS data of all phases, powder diffraction diagrams have been simulated with the CrystalDiffract software using the Crystallographic Information file (CIF) of K3Sb3P2O14 (#1545655) (it should be noted that this structure was later corrected and its true formulation is actually K3Sb3P2O14·1,32 H2O20). The same was done for H3Sb3P2O14, Na3Sb3P2O14, Rb3Sb3P2O14, and Cs3Sb3P2O14 by simply changing the nature of the metallic chemical element in the above CIF file, but keeping all other parameters the same. They are all listed in the SI. A detailed X-ray powder diffraction study of the M3Sb3P2O14·xH2O (M = Li, Na, K, Rb, Cs; x = 0, 5, 6 or 10) compounds, has already been performed by Piffard et al.13
| H3 | H3Sb3P2O14 |
| M3 | M3Sb3P2O14 |
| SAXS | Small angle X-ray scattering |
| WAXS | Wide angle X-ray scattering. |
Footnote |
| † Electronic supplementary information (ESI) available: Fig. SI I: SI1: Photographs of a series of sample tubes with H3Sb3P2O14 weight fraction decreasing from left to right: 1.65, 1.32, 0.83, 0.55, 0.33, and 0.17 w/w% observed (a) in natural light and (b) between crossed polarizers; Fig. SI2: A typical scanning electron microscopy image of H3Sb3P2O14 nanosheets; Fig. SI3: A typical atomic force microscopy image of H3Sb3P2O14 nanosheets with a horizontal cut showing that the nanosheet thickness is ≈1.2 nm; Fig. SI4–7: photographs of 5 series of sample tubes observed in natural light and between crossed polarizers at 4 different concentrations of H3 and different values of x for the cations (Na+, K+, Rb+, Cs+). Fig. SI8: Photographs of five series of sample tubes observed in natural light and between crossed polarizers for H3(1−x)M3xSb3P2O14 nanosheet weight fraction 0.17 w/w%, and variable x values; Fig. SI9–12. SAXS and WAXS patterns of the bottom phase of the samples shown in Fig. SI2 for the [H3(1−x)M3x] systems; SI13: simulated X-ray diffraction patterns for M3Sb3P2O14 (M = Li, Na, Rb, Cs); Fig. SI14: dependence of the lamellar period on the cation exchange rate x for H3(1x)M3xSb3P2O14 (M = Li, Na, Rb, Cs); Fig. SI15: dependence of the pH value on the cation exchange rate x for all series of samples shown in Fig. SI2 (with constant H3Sb3P2O14 weight fraction (1.65 w/w%)); Tables SI1–3: Fast swelling of H3Sb3P2O14 prepared by reverse cationic exchange of M3Sb3P2O14; SI19: Lamellar Lennard-Jones potential; Fig. SI16: SEM-EDX analysis of M3Sb3P2O14 restacked phases. SI II: Animation showing the 2D-corrugated locking mechanism for the restacking of H3(1−x)M3xSb3P2O14. See DOI: https://doi.org/10.1039/d3nr04885c |
| This journal is © The Royal Society of Chemistry 2023 |