Jon A.
Newnham
,
Quinn D.
Gibson‡
,
T. Wesley.
Surta
,
Alexandra
Morscher
,
Troy D.
Manning
,
Luke M.
Daniels
,
John B.
Claridge
and
Matthew J.
Rosseinsky
*
Department of Chemistry, University of Liverpool, Crown Street, Liverpool, L69 7ZD, UK. E-mail: m.j.rosseinsky@liverpool.ac.uk
First published on 30th June 2023
Understanding the structure–property relationships of materials in order to supress thermal conductivity is crucial for developing efficient thermoelectric generators and thermal barrier coatings. Low thermal conductivity materials can often contain a single dominant phonon scattering mechanism. Here, we highlight how combining different structural features into one material can aid in the design and identification of new materials with low thermal conductivities. We synthesise two new mixed-anion materials, Bi8CsO8SeX7 (X = Cl and Br), with low thermal conductivities of 0.27(2) and 0.22(2) W m−1 K−1 respectively, measured along their c-axes at room temperature. The Bi8CsO8SeX7 materials possess a combination of bond strength hierarchies, Cs+ vacancies, and low frequency Cs+ rattling. These different features significantly inhibit phonon transport along different crystallographic directions. Due to sharp bond strength contrast between the van der Waals gaps and [Bi2O2]2+ layers, the Bi8CsO8SeX7 materials exhibit thermal conductivities <50% of the theoretical minimum when measured along the stacking direction. Conversely, the thermal conductivity associated with the ab-plane is reduced by Cs+ rattling when compared to the structurally and compositionally related BiOCl.
Understanding the mechanisms that determine thermal transport in these materials allows us to design and identify new materials with intrinsically low thermal conductivities based on their structure.13,14 These materials often have structural features in common such as bond strength contrast, heavy elements, rattling ions, “liquid-like” movement of ions, and s2 lone pairs.14–20
Combining these strategies to minimise thermal conductivity can be achieved by introducing different structural motifs through the discovery of new materials. For example, Bi4O4SeCl2 is a superlattice of BiOCl and Bi2O2Se, and exhibits thermal transport properties similar to BiOCl along its c-axis, and to Bi2O2Se along its ab-plane.14 The resulting material has an extremely low and highly anisotropic thermal conductivity, highlighting the effectiveness of combining different structural motifs into one material. Mixed anion materials such as Bi2O2Se and BiOCl often display a range of interesting properties including high catalytic activities, high anionic conductivities, and also low thermal conductivities.21–23 This is due to the structural and compositional diversity provided by the use of multiple anions, and the resulting tunability of the materials' properties.24
Despite a range of structural features that can reduce the thermal conductivity of a material, it cannot be arbitrarily low. The amorphous limit describes the minimum thermal conductivity (κmin) of a crystalline material based on two assumptions: (1) the phonon velocities in a material are equal to the Debye velocities, and (2) there is negligible difference between the acoustic and optical phonon velocities.7
BiOCl, Bi2O2Se, and Bi4O4SeCl2 are mixed-anion materials with related structures, compositions, and low thermal conductivities.14 Here, we introduce two new mixed-anion materials into this family, Bi8CsO8SeX7 (X = Cl and Br), and find that they have low thermal conductivities of 0.27(2) and 0.22(2) W m−1 K−1 respectively at room temperature along their stacking (c) axes. These are lower than the theoretical minimum due to a breakdown of the underlying assumptions used to calculate the amorphous limit. Herein, we show that the low thermal conductivities originate from a combination of different structural features including sharp bond strength contrast, Cs+ vacancies, and low frequency Cs+ rattling that interfere with phonon transport.
Bi8CsO8SeCl7 was identified as a target material through simultaneous Bi3+-for-Pb2+ and Se2−-for-Cl− substitutions in Pb0.6Bi3.4Cs0.6O4Cl4.25 Bulk samples of Bi8CsO8SeCl7 were synthesized by hand grinding powders of BiOCl, Bi2O3, Bi2Se3, and CsCl in stoichiometric amounts in an agate pestle and mortar. The resulting mixture was sealed in a 6 mm radius quartz ampule that was evacuated to 10−4 mbar and fired at 700 °C for 12 h using heating and cooling rates of 5 °C min−1. Bi8CsO8SeBr7 was synthesised by grinding BiBr3, Bi2O3, Bi2Se3, and CsBr in stoichiometric amounts in an Ar-filled glovebox (O2 < 0.5 ppm, H2O < 0.5 ppm), and firing them in an evacuated quartz ampule using the same conditions as used to make the chloride. A total of 3 g of each material was synthesised for physical property measurements, and a further 5 g was synthesised for neutron diffraction measurements. We were unable to synthesise the iodide as a pure phase under similar conditions.
The structures of Bi8CsO8SeX7 (X = Cl and Br) were determined using the structure of Bi3.4Pb0.6Cs0.6O4Cl4 as a starting model and refining the atomic positions, displacement parameters, and occupancies of the appropriate sites against neutron diffraction data from banks 1–5 of NOMAD simultaneously.25
The structure of Bi8CsO8SeCl7 that was obtained from the neutron powder diffraction measurements was used as the starting model for the refinements of the VT-PXRD data. Only the background, lattice parameters, atomic positions, and atomic displacement parameters were refined against the data at each temperature. The background was modelled using a Chebyshev polynomial function with twelve parameters. A small, broad background peak at ∼12° 2θ originates from the amorphous SiO2 used to dilute the sample and was fit using a single broad pseudo-Voigt peak that was summed with the background function.
The Einstein and Debye temperatures associated with a particular site can be extracted using the gradient of Uiso(T) using eqn (1) and (2), respectively.27
Uiso = ℏ2T/(mkBθE24π2) | (1) |
Uiso = 3ℏ2T/(mkBθD24π2) | (2) |
The thermal conductivities of Bi8CsO8SeX7 (X = Cl and Br) pellets were measured both parallel and perpendicular to the pressing direction using a Quantum Design Physical Properties Measurement System using the thermal transport option in the temperature range of 2–300 K. Parallel measurements were made using a two-contact method, while perpendicular measurements were made using a four-contact method on a bar cut out of the pellet using a low-speed saw and diamond cutting blade.
Above room temperature, the thermal conductivities were measured by the laser flash method using a Netzsch LFA 457. Measurements were made under dynamic vacuum with a heating rate of 3 K min−1 with 5 min equilibration at each temperature. Thermal diffusivities were obtained by fitting the Cowan model to raw data. The thermal conductivities were then obtained by multiplying the thermal diffusivity, sample density, and heat capacity (assumed to be 3R at high temperatures).
Heat capacity data were modelled using a linear combination of Debye and Einstein terms, and the in-plane thermal conductivities were modelled using the Debye–Callaway model. Full details of modelling the heat capacity and thermal conductivity are given in the ESI,† along with the extracted parameters.
Bi8CsO8SeCl7 has a van der Waals layered structure consisting of two [Bi2O2]2+ fluorite layers that are bridged by a cubic [Cs0.5Cl2]1.5− layer, and are each capped by terminal [Cl]− layer (Fig. 1c). The Cs+ in the [Cs0.5Cl2]1.5− layer occupies the centre of a cubic arrangement of Cl− ions, similar to that of CsCl but with a half-occupied Cs+ site. The selenide anions preferentially occupy the terminal [Cl]− sites over the sites adjacent to Cs+, as the Cl− content refines to 1.000(16) in the [Cs0.5Cl2]1.5− layers, and the anions are fully ordered on this site for Bi8CsO8SeCl7, the occupancy of the Cl− was fixed to 1.0 for the remainder of the refinement. This reflects the stronger bond and more ionic character of Cs–Cl compared to Bi–Se.28 In contrast, there is slight anion mixing on this site in Bi8CsO8SeBr7 as the Br− occupancy refines to 0.980(6), which is likely due to the more similar ionic radii and electronegativities of the Se2− and Br−.
The structure of Bi8CsO8SeCl7 can also be thought of as Cs+ occupying alternate van der Waals (vdW) layers of BiOCl which are then stabilized by substitutions of Cl− for Se2− to balance the charge. The presence of Cs+ alters the layer stacking of BiOCl. Rather than the layers being staggered, they are offset by half a unit cell along the ab-axes to form the cubic environment that the Cs+ can occupy. However, we note that the Cs+ site occupancy cannot be varied by adjusting the Cl−:
Se2− ratio, i.e. Bi8CsxO8SexCl8−x, due to the formation of impurity peaks in the PXRD pattern when x ≠ 1 (Fig. S3†). Due to the structural and compositional similarities of Bi8CsO8SeCl7 to BiOCl, the properties of the two are compared and contrasted throughout the remainder of this paper.
The compositions of the Bi8CsO8SeX7 samples discussed above were confirmed by TEM-EDX. The measured compositions of each material were determined to be Bi8.0(3)Cs0.7(3)Se0.96(6)Cl6.9(5) and Bi8.0(2)Cs0.9(3)Se0.97(4)Br7.6(4), respectively. Oxygen was excluded due to the low accuracy of EDX for measurement of light elements. The measured Bi–Cs–Se–X quaternary plots are given in Fig. S4† and, for both samples, the measured elemental ratios are in good agreement with the refined occupancies from neutron diffraction data and show tight clustering around their nominal compositions.
Homologous analogues of Pb0.6Bi3.4Cs0.6O4Cl4 (the compound that shares its structure with Bi8CsO8SeX7) have been reported, for example, Pb1.5Bi2.5Cs0.5O4Cl3 and Pb0.6Bi1.4Cs0.6O2Cl2.25 However, the respective selenide-containing series, Bi8CsO8Se3Cl3 and Bi4CsO4SeCl3, could not be made via subsolidus routes. The conditions tried are given in the ESI,† and resulting XRD analyses are shown in Fig. S5.†
Both Bi8CsO8SeCl7 and Bi8CsO8SeBr7 are air-stable, and stable up to 700 K under flowing N2 as measured by TGA (Fig. S6†). Above 700 K the material volatilises resulting in a small amount of a poorly crystalline powder. Both materials are insoluble in water and their structures are retained after sonication in water for 24 h, although slight increases in the c-axes of both materials are observed (Table S4†). This increase likely results from the intercalation of water into the van der Waals gaps.
The thermal conductivities of both the Bi8CsO8SeCl7 and Bi8CsO8SeBr7 pellets exhibit a glass-like temperature dependence that plateaus at 0.27(2) and 0.22(2) W m−1 K−1, respectively, when measured parallel to the pressing direction. This linear trend continues up to 600 K when measured by laser flash analysis (Fig. S8†).
The minimum thermal conductivities of both materials were then calculated using the average phonon velocities, which were in turn calculated using the Debye temperatures obtained by fitting the heat capacity data (Fig. 2c–f). This gives a κmin of 0.62 W m−1 K−1 for Bi8CsO8SeCl7, and 0.54 W m−1 K−1 for Bi8CsO8SeBr7 at 300 K. This is particularly unusual as the theoretical minimum is larger than the measured thermal conductivity of each material along the c-axis (i.e. parallel to the pressing direction). This occurs in Bi8CsO8SeX7 because, when measured parallel to the pressing direction, there is considerable bond strength contrast between the vdW gaps and [Bi2O2]2+ layers. This causes a breakdown in the assumptions made in the Debye–Callaway model. Firstly, this implies that there is a reduction in the acoustic phonon velocity relative to the Debye velocity. Secondly, that there is a large difference between the acoustic and optical phonon velocities where the acoustic phonon cut-off frequency (ωc) is lower than the Debye frequency (ωD). The violation of these assumptions that are used to calculate κmin were also observed in BiOCl when measured along the c-axis.14
The thermal transport properties of both Bi8CsO8SeCl7 and Bi8CsO8SeBr7 show significant anisotropy, and the thermal conductivities are larger when measured perpendicular to the pressing direction corresponding to the ab plane of the structures (Fig. 2a and b). Both materials retain the glass-like temperature dependence observed in the out-of-plane data with no low-temperature peak. The thermal conductivity of Bi8CsO8SeCl7 plateaus at 0.83(7) W m−1 K−1 at 160 K, whereas Bi8CsO8SeBr7 plateaus at 0.54(5) W m−1 K−1 at the same temperature. The lack of a crystalline peak in the thermal conductivities suggests that there is significant phonon scattering in the Bi8CsO8SeX7 materials in the low temperature region that supresses the peak in κ typically associated with crystalline materials.29 This typically arises when phonon mean free paths are reduced by scattering mechanisms to distances that are comparable to interatomic distances. The in-plane thermal conductivity of each material can be modelled as they are above κmin. These models are discussed in detail in Section 3.4 and the ESI.†
The uncertainties in these measurements were determined by the PPMS MultiVu software using a combination of the uncertainty associated with the fits of the thermal relaxation profile, and with the uncertainty associated with the corrections applied for radiative heat loss.30 For both materials, there is a somewhat large uncertainty in the perpendicular thermal conductivities due to the larger surface-area to volume ratio in four probe measurement geometry which results in greater radiative heat loss corrections. Despite this, the relative uncertainties observed here are consistent with the relative uncertainties in other materials in the literature when measured by the same method, and it is clear that both materials retain the glass-like temperature dependence observed in the parallel measurements with no low-temperature peak.31
The heat capacities of Bi8CsO8SeX7 were measured and modelled (Fig. 2c–f) using a linear combination of Debye and Einstein terms (see ESI, Table S6†). For both materials, to accurately model the heat capacities at high temperatures (Fig. 2c and d) two Debye temperatures (θD) are required: 180(10) K and 555(10) K for Bi8CsO8SeCl7, and 360(10) K and 510(10) K for Bi8CsO8SeBr7. However, it should be noted that good fits could be obtained with Debye temperatures ±10 K of these values by adjusting the pre-factors of each.
At low temperatures in the Cp/T3(T) plots (Fig. 2e and f), both materials show an asymmetric peak in the measured data which requires three Einstein temperatures (θE) to be modelled. For Bi8CsO8SeCl7 these are 59, 29, and 12 K, while for Bi8CsO8SeBr7, these are 52, 25, and 12 K. Both materials also require a linear (γ) term of 0.33 and 0.53 mJ mol−1 K−2, respectively, to fit the data below 5 K. The γ terms required for Bi8CsO8SeCl7 and Bi8CsO8SeBr7 are of similar magnitude to the γ term required to model the heat capacity of BiOCl, which is 0.39 mJ mol−1 K−2.14 The heat capacities of Bi8CsO8SeX7 cannot be accurately fit with fewer parameters, for example, with only 1 Debye temperature or 2 Einstein temperatures (Fig. S9†).
All of these features are indicative of a loosely bound atom that can disrupt the vibrational modes of its oversized cage.37–39 To investigate the Cs+ rattling hypothesis, we refine the Uiso of each site between 100 K and 500 K (Fig. 3, Table S7†) from synchrotron VT-PXRD data.
![]() | ||
Fig. 3 The refined isotropic displacement parameters (Uiso) of each site in Bi8CsO8SeCl7 as a function of temperature extracted from synchrotron VT-PXRD data. The dashed line plots the Uiso(T) relationship from eqn (1) for a Cs atom with an Einstein temperature of 59 K and a y-intercept of 0.011 Å2. The error bars represent the standard deviations associated with each refinement. |
All Uiso values increase linearly with temperature, however the Uiso(T) of Cs+ exhibits a much steeper gradient compared to the other elements. This shows a much greater displacement of the Cs+ ions relative to the others, and is characteristic of a rattling site or ion.40 The Einstein temperature of the Cs+ can be calculated using the gradient of Uiso(T) using eqn (1) and gives a low θE of 59(6) K. This is identical to one of the Einstein temperatures (59(2) K) extracted via heat capacity measurements, and provides strong evidence for Cs+ rattling within the structure of Bi8CsO8SeCl7, as rattling is a localised vibrational mode.27 The Uiso(T) relationship for a 59 K Einstein oscillator with a mass equal to Cs is overlaid on Fig. 3 as a dashed line. To fit the measured data, a non-zero y-intercept of 0.011 Å2 must be also included. This non-zero intercept is commonly observed for disordered rattling sites, for example in Tl0.22Co4Sb12 and Sr8Ga16Ge30.41,42
We observe that Cl0.75Se0.25 site in Fig. 3 has a slightly less linear relationship of Uiso(T). This site is adjacent to the van der Waals gap and it is therefore reasonable to expect that there might be a greater uncertainty here compared to the other sites as the weak vdW bonding can result in the layers “slipping” past each other resulting in greater disorder and higher uncertainties in the refinement.
The Debye temperatures associated with the other sites can also be calculated using the gradient of Uiso(T) and eqn (2). Using the average gradient associated with the bismuth sites, a calculated Debye temperature of 150(23) K can be obtained which is within error of the lower Debye temperature determined through fitting heat capacity data, 180(10) K (Fig. 2c and e).
While it is likely due to rattling, we rule out other possible origins of the large Cs+ displacement parameters in Bi8CsO8SeCl7. For example, as mobile ions are known to reduce the thermal conductivity of some materials, we investigate the possibility of Cs+ hopping between neighbouring sites.17 To achieve this, the total (electronic + ionic) conductivity was assessed by AC impedance spectroscopy, and the electronic contribution of this conductivity was assessed by DC polarisation measurements (Fig. 4).
The impedance spectrum shows a single semicircle-like relationship between the real and imaginary components with no electrode type feature at low frequencies. The DC polarisation measurements show a single time-independent current when measured at 0.05 V, 0.1 V, 0.3 V and 0.5 V with no tick-like feature at small measurement times (Fig. S10†). The lack of electrode features in AC impedance and tick-like features in DC polarisation indicates that there is no ionic conductivity in Bi8CsO8SeCl7 and that the measured conductivity is solely electronic in nature.
To extract the total (electronic + ionic) conductivity, the impedance spectrum was modelled using ZView2 with a single equivalent circuit containing a resistor and constant phase element (CPE) in parallel.43 The contribution from the electronic conductivity was determined by plotting the measured current from the DC polarisation measurements against the different starting voltages. The conductivities from DC polarisation and AC impedance where determined to be 1.08(2) × 10−6 and 1.4(1) × 10−6 S cm−1 respectively, agreeing with each other and indicating a negligible ionic contribution.
Finally, we rule out the possibility of the large Uiso being as a result of split Cs+ sites as the observed Fourier maps show no evidence of the Cs+ nuclear nor electron densities being elongated along any crystallographic axis (Fig. S11†).42,44 Further, additional Rietveld refinements of the structure with split Cs+ sites always gave negative and unphysical Uiso values for Cs+.
Following confirmation that the large Uiso parameters are a result of Cs+ rattling, we return to the thermal conductivity models discussed earlier. Indeed, for Bi8CsO8SeCl7, in order to obtain any model of the thermal conductivity with a reasonable fit of the data, a 59 K harmonic oscillator must be included in the phonon-lifetime function (see ESI† for details). This 59 K term was extracted from the Cp/T3(T) plots in Fig. 2c and e, and is supported by the Uiso(T) relationship observed for Cs+ in Fig. 3. In the model, this corresponds to phonons with reduced lifetimes in a frequency range surrounding ω59K, the vibrational frequency of the Cs+ atom, hence a reduction in the thermal conductivity. This is similarly true for Bi8CsO8SeBr7, as a 52 K harmonic oscillator must be included in the phonon-lifetime function in order to adequately fit the thermal conductivity. This temperature was also extracted from the measured heat capacity data (Fig. 2d).
The Cs+ vacancies play a significant role in the low thermal conductivity of the material. In the thermal conductivity model for Bi8CsO8SeCl7, the point defect scattering pre-factor (A) is calculated to be 9.1(3) × 10−41 s3, which is an order of magnitude larger than if the Cs+ site is assumed to be fully occupied at 0.62(5) × 10−41 s3. This difference would have a large effect on the thermal conductivity model as, if the Cs+ sites were fully occupied, κ would reach 1.50 W m−1 K−1 at room temperature assuming all of the other scattering pre-factors are unaffected. Whereas, in the model of the measured data, κ reaches 0.92 W m−1 K−1 at room temperature. It is the large contribution from the point defect scattering and resonant scattering processes which dominate at lower temperatures and result in the glass-like thermal conductivity.
There is in fact a smaller mass contrast between disordered anions in Bi8CsO8SeBr7 compared to Bi8CsO8SeCl7, however this has negligible effect on the point defect contribution towards the phonon lifetime as it is dominated by the contributions from the Cs+/vacancy point defects, and from the slower phonon velocities in the bromide.
The lower thermal conductivity in Bi8CsO8SeX7 along the c-axis compared to that along the ab-plane is consistent with the interpretation that the bond strength contrast primarily reduces the thermal conductivity along c by reducing the phonon velocities, whereas the scattering of phonons due to Cs+ rattling affects phonon transport in both the in-plane and c-directions.
Along the c-axis, BiOCl has a lower thermal conductivity (0.15(5) W m−1 K−1) than Bi8CsO8SeCl7 (0.27(2) W m−1 K−1) at room temperature, as filling alternate vdW gaps in BiOCl with Cs+ reduces the amount of bond strength contrast in Bi8CsO8SeCl7.14 In comparison, along the ab-plane, the room temperature thermal conductivity of Bi8CsO8SeCl7 (0.9(2) W m−1 K−1) is lower than that of BiOCl (1.2(1) W m−1 K−1).14 This occurs because, as there are no vdW gaps along the ab-plane in Bi8CsO8SeCl7, phonons are predominantly scattered by the Cs+ rattling which is not present in BiOCl. This highlights the importance of understanding and controlling anisotropy through the combination of unique structural motifs in order to control phonon behaviour within such materials. The thermal conductivity observed in Bi8CsO8SeX7 is also shown in comparison to other low thermal conductivity materials with similar chemistries in Fig. S12.†10,11,14
The main feature that reduces the thermal conductivity in Bi8CsO8SeX7 is the bond strength contrast between the van der Waals gaps and the [Bi2O2]2+ layers. This feature affects phonon transport primarily along the c-direction, leading to very anisotropic phonon propagation. As a result, the thermal conductivity along the c-axis is greatly supressed in both materials (to similar values) as the degree of bond strength contrast is almost identical.
Comparatively, within the ab-plane, the bond strength contrast in Bi8CsO8SeX7 has very little effect on the measured thermal conductivity. This results in greater anisotropy in the chloride compared to the bromide as the higher mass of the Br− anion has a more appreciable effect on the thermal conductivity within the ab-plane. The higher mass of the bromide also results in smaller Debye temperatures and therefore slower phonon velocities (Table S8†).46 The larger mass of Br− also has the effect of increasing the Umklapp scattering pre-factor in the phonon lifetime function for Bi8CsO8SeBr7, as it is proportional to the average atomic mass.47 This has the effect of reducing the thermal conductivity of Bi8CsO8SeBr7 to close to that of κmin at high temperatures.
Bi8CsO8SeCl7 and Bi8CsO8SeBr7 have indirect bandgaps of 1.82(3) eV and 1.88(3) eV respectively, which are too wide for thermoelectric applications. These large band gaps result in the high electrical resistivity that was observed in the DC polarisation measurement of Bi8CsO8SeCl7. However, the band gaps of Bi8CsO8SeX7 (X = Cl, Br) are significantly reduced compared to Bi3.4Pb0.6Cs0.6O4Cl4 (2.97(3) eV) due to the substitution of Se2− for Cl−. Future investigations into a homologous series containing Bi8CsO8SeX7 with differing X−:
Se2− ratios may allow for further optimisation of the band structure and carriers for thermoelectric applications.48,49 The layered nature of the structure also likely make the materials more amenable to combination with other structural motifs to further tune the electronic transport properties, such as the introduction of [BiCuSeO] layers, similar to Bi6Cu2Se4O6.50 Additional routes to doping the materials to increase electronic conductivity and tune the Seebeck coefficient could involve substitution of the Bi3+ or any of the available anion sites.51
For the same reason, the Bi8CsO8SeX7 materials also have narrower optical band gaps compared to the structurally and compositionally similar BiOCl. As a result, they may exhibit potential as photocatalysts where BiOCl is currently used. This could be of particular interest in hydrogen evolution reactions, for example, as Bi8CsO8SeX7 has a similar layered structure and composition to BiOCl, but also a band gap that allows for absorption of visible light, rather than solely in the UV region.21 Similar layered materials have also been investigated for their high dielectric constants and dangling-bond-free surfaces for use in nanoelectronics.52,53
Footnotes |
† Electronic supplementary information (ESI) available: Rietveld fits of neutron diffraction patterns for Bi8CsO8SeX7, PXRD patterns of the additional materials discussed and their extracted lattice parameters, results of thermal and water stability experiments, plots from the DC polarisation measurements at different voltages, and detailed descriptions of the heat capacity and thermal conductivity models can be found in the ESI. The crystal structures of Bi8CsO8SeCl7 and Bi8CsO8SeBr7 are described in the respective crystallographic information files deposited with CSD accession codes 2246936-2246945. The underlying data generated in this study can be found at https://doi.org/10.17638/datacat.liverpool.ac.uk/2156. See DOI: https://doi.org/10.1039/d3ta01630g |
‡ Q. D. Gibson is now based at: Department of Chemistry, University of Aberdeen, Aberdeen AB24 3UE, United Kingdom. |
This journal is © The Royal Society of Chemistry 2023 |