Open Access Article
Erin V.
Phillips
a,
Van Son
Nguyen
b,
Marta
Hatzell
bc and
Carsten
Sievers
*ac
aSchool of Chemistry and Biochemistry, Georgia Institute of Technology, Atlanta, GA 30332, USA. E-mail: carsten.sievers@chbe.gatech.edu
bSchool of Chemical & Biomolecular Engineering, Georgia Institute of Technology, Atlanta, GA 30332, USA
cGeorge W. Woodruff School of Mechanical Engineering, Atlanta, GA 30318, USA
First published on 28th August 2025
Piezoelectric catalysts were synthesized mechanochemically by converting BaCO3 and TiO2 to BaTiO3, Na2CO3 and Nb2O5 to NaNbO3 and Bi2O3 and Fe2O3 to BiFeO3. The catalytic reactivity of BaTiO3, NaNbO3 and BiFeO3 was tested using a mechanocatalytic arylation reaction involving 4-nitrobenzenediazonium tetrafluoroborate. The observed activity in the arylation reaction showed a dependence on the abundance of piezoelectrically active anisotropic phases as measured by the pre-edge intensity in XANES spectra of BaTiO3 and NaNbO3 and distribution of crystalline phases as measured by XRD for BiFeO3. A kinetic analysis showed that the reaction over BaTiO3 was limited by the amount of diazonium salt remaining in the reaction vessel, while the reaction over NaNbO3 and BiFeO3 was limited by the generation of electron hole pairs within the piezoelectric structure. This work shows that mechanochemically produced piezocatalysts have superior structural characteristics such as greater relative abundance of anisotropic phases, higher surface areas and smaller particle sizes that led the mechanochemically produced catalysts to outperform piezoelectric commercial counterparts when tested under the same arylation milling conditions.
The electrical charge transfer that occurs through contact between two materials is known as a triboelectric effect, which has been observed in mechanochemical processes, wherein an electronic charge separation occurs upon mechanical impact.1,16,17 Piezoelectric materials have the ability to form separated charges when a mechanical force is applied because of the polarizability of the piezoelectric unit cell.17–20 Inherent material properties such as the piezoelectric charge coefficient (d), mechanical stress (σ) and the dielectric constant or permittivity (ε) govern the strength of the generated piezoelectric field, E.21,22 A compressed piezo crystal can act as a redox catalyst.16,17,20 For this reason, a mechanochemical environment was chosen for this study, in which ball milling provides direct and repeated impacts to piezoelectric catalysts. Specifically, this work studied the mechanochemical synthesis and reactivity of BaTiO3, NaNbO3, and BiFeO3.
The properties of these materials can be understood in terms of multiferroic behavior, depending on the type of magneto-electric effect that is inherently characteristic of the crystal structure. The simultaneous occurrence of ferroelectricity, ferromagnetism and/or ferroelasticity is referred to as multiferroic behavior and is generated from the inherent distortion of the piezoelectric unit cell. Piezoelectrically active phases include BaTiO3 in tetragonal distortion,23–26 BiFeO3 in a rhombohedral distortion27–30 and NaNbO3 in an orthorhombic distortion.31,32 Having an anisotropic unit cell is a defining characteristic of these structures. Upon application of stress or mechanical impact, the atoms are displaced from their original position, resulting in an overall net electrical charge. Thus, the distortion and elongation of the piezoelectric unit cell creates an electric dipole within the structure. Greater abundance of the desired anisotropic piezoelectric lattice allows for easier electron transfer throughout the structure, as mechanical impact is converted to electrical potential.
However, these asymmetrical distortions are not permanent to the piezoelectric materials and can be altered to a non-active centrosymmetric phase with extended impact in a milling environment. Once the unit cell is centrosymmetric, the cell is no longer polarizable, and the piezo-catalyst is therefore deactivated because it cannot initiate an electron transfer. Depending on the catalyst, a cubic phase can be adopted from milling in the form of a lattice shift, atomic reorganization, or the formation of an entirely new phase. For BaTiO3, the most common form of deactivation is a shift from tetragonal P4mm to cubic Pm3m,24,33–38 where the unit cell remains intact, but the length of the c-axis is shortened, and the distortion of the material is lost. For BiFeO3, Bi2Fe4O9 and Bi25FeO39 (ref. 39–43) are the result of destruction of BiFeO3 and are entirely new structures, which do not present the same strength of piezoelectrical properties, due to increased symmetry. Alternatively, NaNbO3 shifts between different configurations with Pbcm, Pmc21, Ccmm and Pmmn space groups.32,44,45 Many studies have previously examined the synthesis and behavior of piezoelectric materials in a mechanochemical environment with extended milling times. Most techniques utilize planetary mills with zirconia,24,31,33,34,36,44–50 agate35,51 or stainless-steel38,39,42,43,52–54 vessels and milling times up to 30 days. This work, by contrast, examines synthesis using a vibratory mill with a tungsten carbide vessel. Planetary milling typically delivers greater shear forces, whereas vibratory milling offers greater direct impact forces. Many of the articles preceding this work used longer milling times to avoid additional heating.35,41,44,48,52,55–57
Kubota et al.12 first demonstrated the activity of BaTiO3 for mechanoredox arylation and borylation reactions in 2019 using a variety of aryl diazonium salt precursors. TEMPO (2,2,6,6-tetramethylpiperidinoxyl) was also used in this study as a radical scavenger to obtain more complex intermediate products with moderate success. Schumacher et al.58 followed in 2020 with a copper-catalyzed transfer radical cyclization (ATRC) reaction using BaTiO3. Monobromoacetamide was chosen as the alkyl halide, and solid tris(2-pyridylmethyl)amine (TPMA) was chosen as the ligand for the ATRC reaction, while copper triflate, Cu(OTF)2, was the chosen salt; BaTiO3 facilitated the reduction of CuII, which allowed it to be catalytically active for ATRC. Effaty et al.59 and Wang et al.60 followed in 2021 with a multitude of similar basic organic reactions, wherein two substrates were combined to form a singular product in the presence of BaTiO3. In 2022, Wang et al.61 explored the aerobic oxidative coupling of thiols, reaching yields exceeding 99%. In all of these studies, only commercial BaTiO3 was used, and the material-specific properties of the mechanocatalysts were beyond the scope of these studies. On the other hand, piezoelectric materials, such as NaNbO3 and BiFeO3, have been synthesized via milling,32,39,40,42–44,48–51,57,62 but their mechanocatalytic reactivity has not been investigated. In this paper, we report on mechanochemically synthesized piezoelectric materials used in mechanocatalytic reactions, where the synthesized materials exhibited superior performance due to greater surface area, smaller particle sizes, and a higher abundance of desired active phases when compared to commercial catalysts of the same configuration.
:
10 ratio. Passivated Implanted Planar Silicon (PIPS) detector was used to collect fluorescence data. The energy was calibrated using Ti, Nb and Fe foils, measured together with the samples. The edge values were set to 4980 eV (Ti K-edge) for BaTiO3 samples, 19
000 eV (Nb K-edge) for NaNbO3 samples and 7125 eV (Fe K-edge) for BiFeO3 samples. To process the data, the Demeter package was used.63 Between 6 to 9 scans were collected and averaged for each sample to improve data quality.
| BaCO3 + TiO2 → BaTiO3 + CO2 | (EQ1) |
| Na2CO3 + Nb2O5 → 2NaNbO3 + CO2 | (EQ2) |
| Bi2O3 + Fe2O3 → 2BiFeO3 | (EQ3) |
All reactions were conducted using a Retsch MM400 vibrational ball mill with a 25 mL tungsten carbide (WC) milling vessel and 2 × 15 mm WC balls. The precursors were added in equimolar amounts in order to obtain 2 g of the final piezoelectric catalyst. Milling was conducted at 30 Hz from 0.5 to 5 h. Calcination was conducted for BaTiO3 at a ramp of 10 °C per minute for 80 min up to 800 °C and held for 45 minutes before cooling. NaNbO3 and BiFeO3 were also heated at 10 °C per min up to 650 °C and held for 2 h before cooling. To differentiate between the synthesized versus commercial catalysts, catalysts made via milling are denoted as “HM” for “house-made”, followed by milling time (0.5, 1, 2, or 5 h) and “C” to indicate that calcination was completed.
:
15. An Agilent 1260 Infinity HPLC was used for the analysis, equipped with an Agilent Eclipse Plus C18 column. The injected amount was 3 μL at a draw and ejection speed of 200 μL min−1. The flow rate was set at a constant 0.500 mL min−1 flow rate (85% MeOH to 15% H2O), and the program time was set to 7 min. A UV-vis detector was used to analyze product concentration at wavelength of 242 nm.
NaNbO3 was synthesized from NaCO3 and Nb2O5 in equimolar proportions and was found to follow a similar structural evolution to BaTiO3. After 0.5 h of milling, diffractions corresponding to the precursors (NaCO3 and Nb2O5) were no longer detected (Fig. 1C), and new diffraction peaks emerged at 22.5° (002), 32.5° (202), 46.5° (004), 52.5° (204), and 58° (224), confirming the formation of NaNbO3. Rietveld refinement of these peaks was used to determine that the Cmcm phase and the P12/m1 phase were the two main phases present. The phase composition included both an orthorhombic Cmcm structure (lattice parameters: 7.83 Å, 7.84 Å, 7.85 Å) and a monoclinic P12/m1 phase (3.91 Å, 3.91 Å, 3.91 Å).31,32,44,48 The intensity ratio of the peaks at 32.5° and 22.5° allowed differentiation between these phases. In the Ccmm phase, the 32.5° peak was slightly more intense, whereas in P12/m1, the 22.5° peak had a higher intensity. As milling progressed, an increasing fraction of the structure transformed from Ccmm to P12/m1. Since P12/m1 retains an angular distortion (one angle at 90.23°), it is piezoelectrically active, unlike the centrosymmetric cubic phase observed in BaTiO3. Rietveld refinement showed that the relative fraction of P12/m1 increased from 5–7% after 0.5 h to 62% after 5 h, with a corresponding decrease in Ccmm from 93–95% to 38% (Fig. 1D). Further milling induced the transformation of commercial NaNbO3 into the Pmmn:2 phase (lattice parameters: 7.80 Å, 7.80 Å, 15.64 Å), characterized by equal peak intensities at 22.5° and 32.5°. This space group preserved its anisotropy and thus remained piezoelectric. Consequently, NaNbO3 retained its piezoactivity over longer milling time due to the maintained anisotropic phases (Ccmm and P12/m1), with P12/m1 becoming more prevalent.
BiFeO3 was synthesized from Bi2O3 and Fe2O3 under the same milling and calcination conditions as NaNbO3. After 30 min, diffraction peaks at 22° (012), 31.9° (104), 32.2° (110), 39° (006), 39.5° (202), 46° (024), 51.5° (116), 52° (122), 56.5° (018), 57° (214), and 57.3° (030) confirmed the formation of BiFeO3 through Rietveld refinement (Fig. 1E). The dominant lattice was R3c:H (lattice parameters: 5.58 Å, 5.58 Å, 6.39 Å), which is known for its antiferromagnetic properties.57,62,67 Over longer milling time, two secondary phases emerged: the bismuth-rich Bi25FeO39 phase and the iron-rich Bi2Fe4O9 phase. Minor peaks for Bi25FeO39 (I23 space group, lattice parameters: 10.15 Å, 10.15 Å, 10.15 Å) appeared after 30 min and increased with extended milling.39–43,56,68,69. Bi2Fe4O9 (Pbam, lattice parameters: 7.62 Å, 8.11 Å, 5.85 Å) was initially undetectable but became prominent (>1%) after 5 h. Rietveld refinement confirmed a decreasing BiFeO3 abundance, from 94.2% at 0.5 h to 81.5% at 5 h, while Bi25FeO39 increased from 5% to 12%, and Bi2Fe4O9 rose from <1% to 6.5% (Fig. 1F). Although BiFeO3 and Bi2Fe4O9 retained their piezoelectric properties due to their non-centrosymmetric structures, the increasing fraction of Bi25FeO39, which is cubic and centrosymmetric, contributed to reduced piezoelectric activity. The decline in piezoelectric performance over time was correlated with the increasing presence of this inactive phase.
It is important to note that the abundance of phases in BaTiO3, NaNbO3 and BiFeO3 based on Rietveld refinement is reported in relative terms but does not include contribution from possible amorphous phases or phases with concentrations below the detection limit. It is likely that there are other minor phases that form over time with extended milling impact, leading to slight variations in unit cell dimensions. For this reason, only phases with an abundance above 1% were reported. XRD references for BaTiO3 (P4mm, Pm3m), NaNbO3 (Ccmm, P12/m1, Pmmn:2) and BiFeO3 (R3c:H, I23, Pbam) can be found in the SI (Fig. S2, S5 and S7).
In addition to XRD analysis, XAS was conducted to further understand transition states and oxidation states of the central Ti, Nb and Fe atoms, respectively. For BaTiO3, the pre-edge feature at 4970 eV correlates to the excitation from the Ti(1s) to the Ti(3d) unoccupied state (Fig. 2A and B).23,26,70,71 This excitation occurs in tandem with the dipolar transition of the O(2p) orbital to a hybridized Ti(3d) state.30,72,73 This is indicative of ferroelectric off-centering, which is characteristic of strong piezoelectric properties.23,26 A higher intensity of the pre-edge peak indicates a greater concentration of the anisotropic P4mm phase, whereas centrosymmetric phases, such as Pm3m, do not give rise to this characteristic pre-edge peak. BaTiO3 after 1 h of milling had the most intense pre-edge peak, indicating that this catalyst contained a greater abundance of the P4mm lattice,67 rather than the centrosymmetric Pm3m structure.23 The commercial catalyst was the least active and contained the highest amount of Pm3m initially. No significant shift of the XANES edge was observed, and the oxidation state remained constant. Because BaTiO3 was shown to be resistant to the formation of inactive phases and was solely subject to dimensional shifts, BaTiO3 remained the only phase present. Moreover, Ti consistently maintained a +4 oxidation state.
The XAS spectra collected for NaNbO3 (Fig. 2C and D) demonstrated a similar pattern as that in the spectrum of BaTiO3. The spectrum is representative of the hybridization between the Nb(4d) state and the O(2p) state. The conduction band of NaNbO3 is primarily composed of Nb(4d) orbitals while the valence band is mainly composed of O(2p) orbitals.74 The hybridization between the conduction and valence band results in a covalent interaction, which is responsible for atom displacement in NaNbO3.75–77 The pre-edge peak present at 18
990 eV is therefore directly correlated to the presence of a greater relative abundance of the orthorhombic anisotropic phase of NaNbO3. An additional shift of the pre-edge peak to even lower energies, as is observed from NaNbO3 HM1C to HM2C, indicated atom displacement and increasing abundance of the desired NaNbO3 phase.75 Specifically, this shift is indicative of a more highly coordinated Nb–O environment.75 The pre-edge peak for NaNbO3 HM5C nearly aligned with HM1C, suggesting that a similar catalyst activity should be observed. In comparison to the commercial sample, which had the lowest intensity, one would expect to observe reduced catalytic activity. Similar to Ti in BaTiO3, Nb maintained a +5 oxidation state, so the overall shifts of the XANES edge were minimal.
The XANES spectra of BiFeO3 showed minimal changes of the oxidation state of iron with increasing milling time (Fig. 2E and F). With the Fe2O3 precursor, Fe remained as Fe3+ as BiFeO3 was formed. The undesired phases Bi25FeO39 and Bi2Fe4O9 also contained iron as Fe3+. Therefore, no statistically significant shift was observed for the XANES edge position. The pre-edge peak, by contrast, increased in intensity as the realtive abundance of Bi25FeO39 and Bi2Fe4O9 increased. Literature suggests that the pre-edge peak for BiFeO3 is indicative of the electric quadrupole-forbidden transition from the O(1s) level to the Fe(3d) level as hybridization occurs.28,29,78,79 A higher pre-edge intensity correlates to a stronger hybridization, which in this case, may be indicative of increased abundance of the undesired phases and lack of uniformity thoughout the crystalline structure of the sample. BiFeO3 adopts an octahedral hybridization for the central Fe atom, while Bi25FeO39 has a tetragonal arrangement; Bi2Fe4O9 adopts both and is considered to be an isostructural compound, where the Fe atoms is central to two different unit cells.80 In the first cell, Fe1O4, Fe is surreounded by four O atoms in a tetrahedral arrangement, while the second cell, Fe2O6, is octohedral. The increase in pre-edge peak intensity suggested increased abundance of nonuniform hybridized orbitals. Thus, the XANES results found in this study correlated to intermediate structures as the BiFeO3 phase was converted to Bi25FeO39 and Bi2Fe4O9.
![]() | ||
| Fig. 3 (A) Surface area based on N2 physisorption results for the three catalysts. (B) Average particle size for BaTiO3, NaNbO3 and BiFeO3 based on TEM images. | ||
The particle sizes of the mechanochemically synthesized catalysts as determined by TEM image analysis averaged 250 nm or smaller in diameter, while that of the commercial catalysts averaged more than 400 nm in diameter (Fig. 3B and Fig. S14–S16). With extended milling, the particle size of BaTiO3 increased from 36 nm after 1 h to 178 nm after 5 h. Commercial BaTiO3 had an average diameter of 482 nm. For NaNbO3 the initial size after 1 h was 54 nm, which increased to 75 nm after 5 h. Commercial NaNbO3 had an average diameter of 426 nm. The average particle size of BiFeO3 remained steady between 220–250 nm from 1 h to 5 h of milling, indicating that extended milling had a minimal impact. Results for pore diameter and pore volume can be found in the SI (Fig. S8) while histograms of all particle sizes can be found in Fig. S14–S16.
The TEM micrographs of mechanochemically synthesized BaTiO3 and NaNbO3 contained areas in which parallel lattice plains were clearly visible, whereas commercial BaTiO3 and NaNbO3 contained less defined features (Fig. 4). The observable crystalline domains for the milled BaTiO3 and NaNbO3 imply strong crystallinity, which is expected to correlate to greater conductivity. This conductivity allowed for a more efficient transfer of electricity throughout the catalyst structure, due to greater ordering, resulting in enhanced electron movement. Additionally, there were distinct difference in particle size between the commercial and milled BaTiO3 and NaNbO3 (Fig. S14 and S15). Commercial NaNbO3 had large particles spanning from 100 nm up to 1600 nm in diameter, while the milled NaNbO3 showed a greater consistency of square-like particles ranging from roughly 25–150 nm in diameter. Similarly, commercial BaTiO3 consisted of particles ranging from 100 nm to 1100 nm, while the mechanochemically synthesized particles did not exceed 200 nm until after 2 h of milling. These histograms (Fig. S14–S16) demonstrate that mechanochemical synthesis of the piezoelectric catalysts results in more uniform particle sizes and overall higher surface area, leading to a greater number of exposed active sites.
![]() | ||
| Fig. 4 TEM images for commercial BaTiO3 (A) at a 50 nm scale and HM1C BaTiO3 (B) at a 20 nm scale. TEM images for commercial NaNbO3 (C) at a 200 nm scale and HM2C NaNbO3 (D) at a 50 nm scale. | ||
TEM images of all mechanochemically synthesized samples of BaTiO3, NaNbO3 and BiFeO3 can be found in Fig. S11–S13. For BiFeO3 HM1C parallel lattice plains were clearly visible throughout the particle structure (Fig. S13B), congruent with observations made for BaTiO3 and NaNbO3 in Fig. 4. Additionally, SEM images depicting the morphological differences and comparative surface roughness between the commercial and mechanochemically synthesized BaTiO3 and NaNbO3 catalysts are shown in Fig. S9 and S10.
![]() | ||
| Scheme 1 Arylation of 4-nitrobenzenediazonium tetrafluoroborate and furan, resulting in 2-(4-nitrophenyl) furan as the main product. | ||
Control experiments included milling 4-nitrobenzenediazonium tetrafluoroborate alone (Fig. S23) and milling the salt with furan (Fig. 5). The first experiment showed that when the salt was milled alone, only 70% was recollected after milling. This was due to salt decomposition during milling, which was visible upon opening the vessel after the reaction completed. When the salt was milled with furan only, roughly 50% of the salt remained and 12% of 2-(4-nitrophenyl) furan was produced. Once the seal of the vessel was broken, a pressure release ensued as particle matter escaped the vessel and dissipated. The pressure release indicated that the salt partially decomposed to gaseous components, but because this reaction was conducted in a closed milling vessel, the gaseous products could not be collected for quantitative analysis.
The low conversion of the diazonium salt and 12% yield of 2-(4-nitrophenyl) furan demonstrated that a catalyst was necessary to facilitate the radical mediated arylation reaction. The BaTiO3 samples that were milled for 1 h during synthesis produced the highest yield of 2-(4-nitrophenyl) furan at 40% (Fig. 6). Lower yields of 33% and 31% were observed for the catalysts milled for 2 and 5 h during their synthesis, respectively. The commercial catalyst produced 2-(4-nitrophenyl) furan with a yield of 25%, which was lower than all mechanochemically synthesized samples of BaTiO3. This activity correlated to XAS findings specifically, where BaTiO3 prepared via 1 h of synthesis milling had the strongest pre-edge feature, and therefore the highest relative abundance of the P4mm anisotropic phase.
![]() | ||
| Fig. 6 2-(4-Nitrophenyl) furan yields from the arylation of 4-nitrobenzenediazonium tetrafluoroborate with furan to over BaTiO3, NaNbO3 and BiFeO3 (C). | ||
Using NaNbO3, the conversion of 4-nitrobenzenediazonium tetrafluoroborate increased steadily with increased synthesis milling time from 0.5 h (38% yield 2-(4-nitrophenyl) furan) to 2 h (44% yield) followed by a decrease in activity for the sample milled for 5 h (37% yield) (Fig. 6). The low variation in yields (only 7%) are in line with the XRD analysis that showed that NaNbO3 maintained a relatively stable structure and piezoelectric potential with extended synthesis milling. The fresh commercial catalyst gave a 19% yield, while the calcined commercial catalyst yielded 31%. As shown by XRD (Fig. S4), the commercial NaNbO3 was not yet fully formed and required additional heating and milling to reach a pure NaNbO3 phase with no Nb2O5 present.
The HM0.5C catalyst was most productive for BiFeO3, which produced a 32% yield, followed by a decrease for the samples with synthesis milling times of 1 h (28%), 2 h (25%), and 5 h (24%), respectively (Fig. 6). This trend is in agreement with the observation of increasing amounts of Bi25FeO40 and Bi2Fe4O9 by XRD, indicating that 0.5 h of synthesis milling produced the purest and therefore most active catalyst. Conversion over all catalysts can be found in Section S2.3 (Fig. S22), while 1H NMR spectra and product analysis can be found in Section S2.2 (Fig. S17–S21).
![]() | ||
| Fig. 7 Changes in 2-(4-nitrophenyl) furan yield for BaTiO3 catalysts compared to changes in normalized XAS pre-edge peak intensity. | ||
By comparison, the surface area increased throughout 5 h of milling, which indicated that size reduction of particles occurred, but this was coupled with the formation of centrosymmetric cubic unit cells and, thus, catalyst deactivation. The mechanochemically synthesized BaTiO3 surpassed the yield of the arylation product over the commercial sample by 20%, likely because the synthesized BaTiO3 had well defined lattice plains, as observed by TEM in Fig. 4. The commercial sample lacked observable crystalline domains and was therefore less conductive and crystalline when compared to the mechanochemically synthesized sample. The commercial sample also had half the surface area as the mechanochemically milled particles, suggesting fewer active sites were initially accessible on the larger particles. Overall, the characterization and test reaction confirm that the BaTiO3 sample that was mechanochemically synthesized for 1 h was the most effective out of all BaTiO3 samples because it had the greatest abundance of the desired P4mm anisotropic phase.
NaNbO3 demonstrated very similar behavior to BaTiO3, and reactivity for this piezoelectric material also scaled with the pre-edge peak intensity of the sample, correlating to greater hybridization of the Nb(4d) and O(2p) orbitals and a greater abundance of the orthorhombic Ccmm phase (Fig. 8). Because XRD confirmed that both Ccmm and P12/m1 phases were present and anisotropic, decreases in the catalytic activity of this catalyst could not primarily be attributed to changes in crystallinity. The shift from Ccmm to P12/m1 coupled with a reduction in surface area, presumably by sintering, resulted in less active catalysts with excessive milling time during synthesis. This shift in crystal phase was also identified by XAS (Fig. 2C and D), and a significant decrease in peak intensity was observed from 2 h of synthesis milling to 5 h. The HM2C sample had an intensity of 0.201, which fell to 0.175 for HM5C; the intensity for HM1C fell in between these two values at 0.179. As shown by Fig. 8, these values strongly correlated to the yields obtained for 2-(4-nitrophenyl) furan. The analysis of surface area gave greater insight into the reactivity observed for the arylation reaction, where NaNbO3 synthesized for 1 h and 2 h were the most highly reactive catalysts and had surface areas of 10.4 and 10.5 m2 g−1 respectively. This resulted in yields of 41% and 44% 2-(4-nitrophenyl) furan. After 5 h, particle sintering began to occur, and the surface area fell to 3.4 m2 g−1, followed by a decrease in yield to 37%. By contrast, the commercial NaNbO3 had the lowest initial surface area and pre-edge peak intensity at 0.9 m2 g−1 and 0.162 respectively and achieved a 19% yield of 2-(4-nitrophenyl) furan.
![]() | ||
| Fig. 8 Changes in 2-(4-nitrophenyl) furan yield for NaNbO3 catalysts compared to changes in normalized XAS pre-edge peak intensity. | ||
Like BaTiO3, NaNbO3 outperformed its commercial counterpart, doubling the yield. As shown in Fig. S4, the untreated catalyst did not have a fully formed crystal structure, so calcination was performed in an attempt to remedy this issue. After heating, characteristic peaks for NaNbO3 became more prominent, but it was not until the catalyst was milled that the Pmmn:2 crystal structure was identifiable. The larger particle size of the commercial NaNbO3 catalyst was also confirmed through TEM imaging (Fig. 3 and 4). Physisorption showed that the surface area of the commercial catalyst was one tenth that of the mechanochemically synthesized catalyst, and TEM images demonstrated that the particle size (up to 1600 nm diameter) was the largest of all the analyzed piezoelectric catalysts. Fig. 4 also depicts the uniform rounded cubic structure of the mechanochemically synthesized particles, whereas the commercial NaNbO3 sample had much larger particles that lacked observable lattice plains, similar to BaTiO3. Calcination improved the yield of 2-(4-nitrophenyl) furan over commercial NaNbO3 by roughly 12%, but it still underperformed when compared to the worst performing mechanochemically synthesized NaNbO3 catalyst.
For BiFeO3, the reactivity of the catalyst directly correlated to the increase in the abundance of inactive phases as observed by the XRD results (Fig. 9). The undesired Bi25FeO40 and Bi2Fe4O9 phases were already apparent after the first 30 min of synthesis milling. As the abundance of these phases increased, conversion fell, and yields decreased for the arylation reaction from 32% to 24%. While extended milling had a significant impact on the crystallinity of BiFeO3, changes in the surface area and particle size were negligible.
![]() | ||
| Fig. 9 Changes in 2-(4-nitrophenyl) furan yield for BiFeO3 catalysts compared to increasing composition of the Bi4Fe2O9 and Bi25FeO39 phases as determined by XRD. | ||
| ehp → e− + h+ | (R1) |
![]() | (R2) |
![]() | (R3) |
![]() | (R4) |
![]() | (R5) |
![]() | (R6) |
The conversion of 4-nitrobenzenediazonium tetrafluoroborate as well as the rate of formation and the yield of 2-(4-nitrophenyl) furan and 4,4′-dinitrobiphenyl observed for each catalyst are plotted against time in Fig. 10.
The conversion of 4-nitrobenzenediazonium tetrafluoroborate reached approximately 90% over all three catalysts after 60 minutes of milling (Fig. 10A). The measured rate of formation for 2-(4-nitrophenyl) furan for all catalysts are represented in Fig. 10C. Specifically, the formation of 2-(4-nitrophenyl) furan was shown to follow a first order in 4-nitrobenzenediazonium tetrafluoroborate when applied to BaTiO3, whereas NaNbO3 and BiFeO3 followed zero-order and ½-order in 4-nitrobenzenediazonium tetrafluoroborate, respectively. Similarly, the measured rate of formation of 4,4′-dinitrobiphenyl in Fig. 10E followed first-order kinetics observed for BaTiO3 and BiFeO3, whereas NaNbO3 exhibited a zero-order in 4-nitrobenzenediazonium tetrafluoroborate for this reaction. In addition, the kinetics provided insight into the limiting factors of the arylation reaction of 2-(4-nitrophenyl) furan with 4-nitrobenzenediazonium tetrafluoroborate. To test potential mechanisms, rate equations were derived based on various assumptions. The protonation step (eqn (R5)) was assumed to be in quasi-equilibrium, with equilibrium K5 (see eqn (SEQ1)). Due to the high reactivity of radicals, the pseudo-steady-state hypothesis (PSSH) was applied to assume that the concentration of the nitrophenyl ((O2NPh)˙) and the 2-(4-nitrophenyl) furan radicals (O2NPhC4H4O˙) remained constant (see eqn (SEQ2)). With furan in large excess, the overall rate expression was assumed to be independent of the furan concentration, i.e., quasi zero-order (see eqn (SEQ3)).
For the formation of 2-(4-nitrophenyl) furan, the reaction rate (rO2NPhC4H4O) and reaction order differed among catalysts due to variations in piezoelectric properties and the availability of charge carriers. BaTiO3 exhibited a first-order dependence on the diazonium salt. Under the pseudo-steady-state hypothesis (PSSH), the concentration of the nitrophenyl radical intermediate was shown to be approximated as a function of the diazonium salt and electron concentration (see eqn (SEQ6)). The latter was found to be governed by the strength of the piezoelectric field E, which is determined by intrinsic material properties specifically, the piezoelectric charge coefficient d, the mechanical stress σ, and the dielectric constant (or relative permittivity) ε
.21,22 The resulting electron–hole pair concentration ([e−]/[h+]) was treated as a material-specific parameter 
, which encapsulates the effective concentration of charge carriers generated under a given field strength. Substituting this into the rate law yields a first-order dependence on the diazonium salt (rO2NPhC4H4O,BaTiO3 = kC,BaTiO3,app[(O2NPhN2)+BF4−]), in which the apparent rate constant can be expressed as
(see eqn (SEQ10)).
The rate of formation of 2-(4-nitrophenyl) furan followed zero-order kinetics for NaNbO3 in 4-nitrobenzenediazonium tetrafluoroborate (rO2NPhC4H4O,NaNbO3 = kC,NaNbO3,app), indicating that the reaction rate was independent of the diazonium salt concentration. This behavior is attributed to a rate-limiting step governed by the generation of electron–hole pairs in the piezocatalyst, which occurs more slowly than subsequent transformation steps (see eqn (SEQ12)). As a result, the apparent rate constant kC,NaNbO3,app is determined primarily by the intrinsic ability of NaNbO3 to generate charge carriers under mechanical activation
.
BiFeO3 demonstrated a half-order in 4-nitrobenzenediazonium tetrafluoroborate (rO2NPhC4H4O,BiFeO3 = kC,BiFeO3,app[(O2NPhN2)+BF4−]1/2), (see eqn (SEQ17)). The radical concentration [(O2NPh)˙] was derived using the pseudo-steady-state hypothesis (PSSH) and shown to be proportional to [(O2NPhN2)+BF4−]1/2 (see eqn (SEQ15)). Applying PSSH to the concentration of the 2-(4-nitrophenyl) furan radical species [O2NPhC4H4O˙] led to the reaction rate follows a half-order dependence on the diazonium salt concentration (see eqn (SEQ16) and (SEQ17)). This fractional order reflects the enhanced charge carrier availability due to its relatively high charge density. The lower reaction rate over BiFeO3 compared to other materials, suggested that the piezocatalytic activity of BiFeO3 tends to be limited because a significant fraction of it is present in inactive phases such as Bi25FeO40 and Bi2Fe4O9 (see Section 4.1).
The formation of 4,4′-dinitrobiphenyl over BaTiO3 followed first-order kinetics (r(O2NPh)2,BaTiO3 = kE,BaTiO3,app[(O2NPhN2)+BF4−]) (see eqn (SEQ11)). This first-order dependence appears as the [(O2NPh)˙] remains low ([(O2NPh)˙] ≈0), allowing its quadratic term in the rate expression to be approximately close to the value of the first-order term ([(O2NPh)˙] ≈ [(O2NPh)˙]2) (see eqn (SEQ11)). Since the rate of formation of 4,4′-dinitrobiphenyl was proportional to [(O2NPh)˙]2, substituting this expression yields an overall first-order dependence on the diazonium salt (see eqn (SEQ13)). NaNbO3, in contrast, displayed zero-order kinetics (r(O2NPh)2,NaNbO3 = kE,NaNbO3,app) because the rate was limited by the availability of charge carriers (see eqn (SEQ13)). BiFeO3 exhibited first-order kinetics (r(O2NPh)2,BiFeO3 = kE,BiFeO3,app[(O2NPhN2)+BF4−]), indicating that the rate of dimer formation scaled with the concentration of the nitrophenyl radicals (see eqn (SEQ18)). The apparent reaction rate constants for the formation of 2-(4-nitrophenyl) furan (kC,app) and of 4,4′-dinitrobiphenyl (kE,app) for all cases are provided in Fig. 10C and E. The calculated yields of 2-(4-nitrophenyl) furan and 4,4′-dinitrobiphenyl aligned well with measured yields, demonstrating an accurate estimation of the reaction rate across all three catalysts (Fig. 10B and D).
The present results show that several descriptors affect the performance of the piezoelectric catalysts in this study, which plays into the reaction rates by influencing the concentration of electrons available for the piezocatalytic activity. While a sufficient surface area is needed for reactants to interact with the catalyst, the catalyst also needs to be able to separate charges, which depends on the presence of piezoelectric phases and properties of these phases, such as piezoelectric charge coefficient, electromechanical coupling factor, and the dielectric constant.81 In the present case, the formation of 2-(4-nitrophenyl) furan as well as 4,4′-dinitrobiphenyl was favored over catalysts with high surface area and abundance of piezoelectric phases. NaNbO3 thus had the highest yield for 2-(4-nitrophenyl) production (44%), correlating with its stable structure and high surface area with milling time. Therefore, it is essential to consider both microscopic properties (e.g., coupling factor, crystal structure) and macroscopic characteristics (e.g., surface area) to ensure optimal catalytic performance. The interplay between these factors can significantly influence reaction kinetics, leading to more complex mechanistic behavior. While the radical nature of the aryl coupling mechanism has not been directly demonstrated in this study, our kinetic observations are consistent with the mechanism of radical-mediated arylation involving aryl diazonium salts reported by Kubota et al.12 The use of radical traps such as TEMPO is a well-established approach for probing radical mechanisms,82 and the results of the Kubota study provide strong support for the intermediacy of aryl radicals under related conditions.12
Overall, piezocatalytic activity appears to be governed by the surface area of anisotropic phases and their specific activity. The latter can be expressed as the overall piezoelectric field strength of each catalyst. However, field strength cannot be directly measured during the mechanochemical process in this work. The abundance of these phases was observed in a different way for each catalyst. The activity of BaTiO3 was directly proportional to XANES pre-edge intensity, which was indicative of the abundance of the anisotropic P4mm tetragonal phase. The activity of NaNbO3 also scaled with the abundance of anisotropic phases, but it also increased with increasing surface area, demonstrating the need for small piezoelectric particles. By contrast, the activity of BiFeO3 decreased as the abundance of non-piezoelectric Bi4Fe2O9 and Bi25FeO39 phases increased. These findings underscore the necessity of considering multiple material properties in catalyst design to optimize reaction kinetics and product selectivity. This work provides guidance for designing mechanocatalytic processes, wherein the piezoelectric catalyst can be efficiently activated via mechanical impacts. The energy provided by milling coupled with optimal crystallographic features for these materials created active species for radical reactions.
Additional data supporting this article have been included as part of the SI. Supplementary information: Analytical data including XRD, N2 physisorption analysis of pore size and volume, SEM and TEM images with histograms, a schematic for side product formation of the arylation reaction, derivation of kinetic rate equations, proton 1H NMR spectra, conversion data and FTIR spectra. See DOI: https://doi.org/10.1039/d5mr00062a.
| This journal is © The Royal Society of Chemistry 2025 |