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

Calorimetric determination of the heat capacity function and absolute entropy of yttrium borohydride (Y(BH4)3) mechanochemically prepared

Konrad Burkmann a, Franziska Habermann a, Bianca Störr a, Jürgen Seidel a, Roman Gumeniuk b, Klaus Bohmhammel a and Florian Mertens *ac
aInstitut für Physikalische Chemie, TU Bergakademie Freiberg, Leipziger Straße 29, 09599 Freiberg, Germany. E-mail: florian.mertens@chemie.tu-freiberg.de; Tel: +49-3731-393737
bInstitut für Experimentelle Physik, TU Bergakademie Freiberg, Leipziger Straße 23, 09599 Freiberg, Germany
cZentrum für Effiziente Hochtemperatur-Stoffwandlung, TU Bergakademie Freiberg, Winklerstraße 5, 09599 Freiberg, Germany

Received 25th October 2024 , Accepted 28th April 2025

First published on 6th May 2025


Abstract

Over the last two decades, complex metal hydrides have attracted attention in various fields such as chemical hydrogen storage, solid state electrolytes and superconductivity. In this context, we determined thermodynamic properties of the complex metal hydride Y(BH4)3. Yttrium borohydride was prepared by solid state metathesis yielding a mixture containing three equivalents lithium chloride beside the boranate. The benefit of the mechanochemical preparation procedure compared to the classical wet chemical one is to avoid the desolvation step which can lead to a partial decomposition of the hydride. The heat capacity of the compound as a function of the temperature covering a temperature range from 2 K to 370 K was determined using two different calorimetric techniques. Based on the heat capacity data, the standard entropy at 298.15 K was obtained. From the evaluation of the low temperature heat capacity region the Sommerfeld coefficient and the Debye temperature were derived.


Introduction

Even after many years of intensive research, complex hydrides continue to be the focus of current research in the fields of chemical hydrogen storage,1–11 solid state electrolytes12,13 and superconductivity.14 Yttrium borohydride, Y(BH4)3, is a salt like borohydride with a theoretical total hydrogen content of about 9.1% and a decomposition temperature of 460 K2. Although reversibility has not been demonstrated yet for this material,15 the compound may nevertheless be of interest because the dehydrogenation reaction is endothermic and, thus, the material may be suitable for rehydrogenation using a catalyst. It may also potentially be used as a component for modifying hydride mixtures by applying the concept of thermodynamic tuning to achieve reversibility for the overall reaction.16,17 Therefore, reliable thermodynamic data are needed to predict the decomposition behaviour (e.g. decomposition temperature and products, equilibrium hydrogen pressure) of a given hydride mixture containing Y(BH4)3.

Several synthesis routes have been described for the synthesis of the α polymorph of Y(BH4)3 in the literature.2,7 Nevertheless, solvent mediated or mechanochemically driven metathesis reactions without solvents using a ball mill are the two most common methods to synthesise Y(BH4)3 from YCl3 and an alkali metal borohydride (LiBH4 or NaBH4).2,7 The reaction between yttrium chloride and lithium borohydride is given by eqn (1):

 
3LiBH4 + YCl3 → Y(BH4)3 + 3LiCl(1)

Jaroń and Grochala showed that the ball milling reaction did not take place by using sodium borohydride.18 Several authors used ball milling for the synthesis of Y(BH4)3 as well19–26 resulting in mixtures consisting of Y(BH4)3 and LiCl. For the solvent mediated metathesis DMS,27 THF18,28–30 or diethyl ether15,31 were used. Besides the formation of stable solvent adducts, another problem is given by the solubility of the by-product LiCl in the used solvent,15,32,33 except for DMS29 in which LiCl is insoluble. Therefore, when using THF or diethyl ether as a reaction medium, it is advised to extract the product mixture with DMS to obtain a pure DMS-solvent adduct29,31 or to use NaBH4 instead of LiBH4 as a reactant for the synthesis,28 because NaCl is insoluble in diethyl ether and THF.33 Certainly, the extraction step can also be applied when using the mechanochemical route. Any method that uses solvents must be complemented by a thermal desolvation step to obtain the pure Y(BH4)3.29 However, very often uncontrolled hydrogen desorption occurs during the decomposition of the solvent complexes. In addition, the solvent itself can decompose during desolvation and thus contaminate the sample.

Up to our best knowledge, only one publication addresses the determination of thermodynamic quantities of Y(BH4)3 so far. Lee et al. calculated the reaction enthalpies and reaction entropies for the decomposition of Y(BH4)3 into several yttrium borides using the Quantum-ESPRESSO density functional theory34 from which we derived the enthalpies of formation and the absolute entropies of α-Y(BH4)3 and β-Y(BH4)3, resulting in the following values: ΔFH (298.15 K) = −409.8 kJ mol−1 and S (298.15 K) = 149.4 J mol−1 K−1 for α-Y(BH4)3 and ΔFH (298.15 K) = −405.9 kJ mol−1 and S (298.15 K) = 153.4 J mol−1 K−1 for the high temperature β polymorph of Y(BH4)3. The experimental determined temperature of the phase transition is about 454 K.25,35,36 Nevertheless, no experimental values are available for both thermodynamic quantities.

The aim of our work is to determine experimentally the standard entropy of α-Y(BH4)3 using its measured heat capacity function. We used a solid state metathesis reaction leaving out the solvent extraction step that is often applied in complex metal hydride syntheses.7,29 The latter to avoid decomposition and contamination with the solvent of the hydride halide mixtures (Y(BH4)3 + 3LiCl). To the best of our knowledge, this procedure has not been shown to be applicable to boranates and our study is the first to confirm that, if the composition of the mechanochemically produced mixtures is known, accurate thermodynamic data for the boranate can be obtained without separating the by-product prior to the respective calorimetric measurements. In future, this will further establish mechanochemistry as a simple and environmentally friendly method for the production of various hydrides, if their thermodynamic data are to be determined. This procedure is justified, because no double salts, like in the case of lanthanum borohydride forming LaLi(BH4)3Cl,37,38 are formed and therefore the by-product LiCl can be assumed to be inert.

Experimental section

Materials

All handling and manipulation of the chemicals and the sample preparation for the measurements were performed in an argon (Nippon Gases, 5N) filled glove box with a gas circulation purifying system (H2O and O2 < 0.1 ppm) from MBRAUN.

Lithium borohydride (LiBH4, Chemetall) and yttrium trichloride (YCl3, Alfa Aesar, ultra dry, 99.99%) were used as received. Sapphire, (aluminium oxide, Al2O3) was used as reference material for the heat capacity measurements. A 25 μm thick copper foil (Cu, Alfa Aesar, 99.999%) was used to fold copper crucibles as containers for the samples used in the PPMS measurement.

Characterisation techniques

Powder X-ray diffraction (PXRD). The measurements were performed with a D2 Phaser from Bruker (5° to 85° 2θ, 0.05° 2θ steps, dwell time 1 s, 300 W X-ray power, λ(Kα) = 1.54184 nm) with a LYNXEYE detector. The powders were ground with a mortar and pestle and placed on a steel plate. To prevent oxidation of the samples during the measurement, the plate was covered with polyethylene foil.
Calorimetric heat capacity measurements. The heat capacities were measured in a temperature range from 2 K to 380 K based on the method described by ref. 4, 5, 8 and 9. The measurements in the low temperature region from 2 K to 300 K were conducted with a Physical Property Measurement System (PPMS, DynaCool-12, Quantum Design, USA) equipped with the heat capacity option, which is based on a relaxation technique described elsewhere.39,40 The preparation of small copper crucibles containing the powdered sample is explained by ref. 41–43. To ensure a good thermal contact between the copper crucibles and the sample platform, Apiezon N grease was used, possessing besides a good thermal conductivity also a low heat capacity.43–45 The PPMS software automatically computes the heat capacity, by subtracting the values of the addenda measurement (sample platform and grease) from the measured data of the sample (sample and crucible, sample platform and grease).42 The heat capacity of the sample CP,sample is then calculated as the difference between the total measured specific heat CP,total and the specific heat of the copper crucible CP,crucible, which is calculated from literature heat capacity data,46 weighted according to their mass fractions.
 
image file: d4mr00124a-t1.tif(2)

Furthermore, the same approach is used to obtain the specific heat of the Y(BH4)3 from the measured sample mixture of Y(BH4)3 + 3LiCl by subtracting the specific heat of LiCl from it using the data published by Shirley47 and Moyer48 for LiCl for temperatures lower than 300 K and from NIST49 for temperatures higher than 300 K. The heat capacity measurements at temperatures higher than 280 K were performed in a DSC 111 from SETARAM, France, using the instrument software (Calisto, AKTS/SETARAM) to conduct the data evaluation and were inspired by the method already described elsewhere.4–6,8,9,42,50–54 In detail, about 100 mg of the sample were filled into standard stainless steel sample vessels. Using a nickel sealing the vessels were crimped tightly. The nickel sealings and the vessels used for the sample, the reference and the blank were brought to an identical mass by hand polishing to avoid variations in the blank effects. A CP-by-step method55 based on four temperature steps (5 K, 3 K min−1) each followed by an isothermal period of 30 min duration was conducted in the temperature range from 10 °C to 30 °C. After that, another step programme based on eight steps (10 K, 3 K min−1) each also followed by an isothermal period of 30 min duration in the temperature range from 30 °C to 100 °C was performed. This procedure was used for the measurements of the sample, the reference (granular sapphire for the verification of the overall accuracy of the experiments) and the blank. Using the recommended reference heat capacity of sapphire published by Della Gatta et al.,56 the specific heat of the sample were calculated viaeqn (3).

 
image file: d4mr00124a-t2.tif(3)
[C with combining macron]P,sample denotes the specific heat of the sample at the mean temperature of the ramp. The begin and end of the heat flow peaks [Q with combining dot above] are implied by the times ti and ti+1. The masses of the sample and reference material are given by msample and mref, respectively. Finally, [C with combining macron]P,ref is used as the specific heat of the reference material at the mean temperature of the ramp. The molar heat capacity values can be derived by an multiplication with the molar mass of the Y(BH4)3 (MY(BH4)3).

Synthesis

Lithium borohydride and yttrium trichloride were ball milled in stoichiometric amounts (see reaction (1)) essentially following the published method from Ley et al.27 A Fritsch Pulverisette 6 planetary ball mill equipped with a 80 mL-tungsten carbide (WC) crucible and three WC balls with an outer diameter of about 10 mm and 7.1 g of weight each were used. The milling programme applied consisted of 2 min milling followed by 2 min pause. This procedure was repeated for 179 times (total milling time of 6 h). A ball-to-powder-ratio of about 35[thin space (1/6-em)]:[thin space (1/6-em)]1 and rotational speed of 400 rpm were applied. The material was milled under protective atmosphere which was achieved by flooding the crucible with 1 bar of argon. After milling, the product mixture was removed from the crucible and stored for further investigations in the glove box.

Results and discussion

Phase purity of the synthesised samples using PXRD

From the absence of LiBH4 and YCl3 reflections in the diffractogram of the product mixture (Y(BH4)3 + 3LiCl), a complete conversion of the reactants can be inferred (Fig. 1).
image file: d4mr00124a-f1.tif
Fig. 1 PXRD of the product mixture after the metathesis reaction according to eqn (1). The reflections of the assigned phases were taken from the ICSD.57

Furthermore, there are no reflections of possible decomposition products such as YH2, YH3, B or other boron containing species visible indicating that there was no or negligible decomposition during milling, so the used procedure appears appropriate for the preparation of the compound.

Heat capacity function of Y(BH4)3

The experimental determination of the heat capacity of Y(BH4)3 was carried out over a broad temperature range from about 2 K to 370 K. To obtain the specific heat capacity values of Y(BH4)3 listed in Table 1 from the experimental values, the heat capacity of copper (crucible) and LiCl (by-product) were subtracted according to their mass fraction in the sample composition. The results of that calculation are presented in Fig. 2.
Table 1 Experimental heat capacity values for Y(BH4)3 (M = 133.4341 g mol−1) at p = 100 kPa
T [K] C P [J mol−1 K−1] T [K] C P [J mol−1 K−1] T [K] C P [J mol−1 K−1]
2.0606 0.014475 8.9878 0.91862 16.132 3.9476
2.1413 0.015631 9.0881 0.95080 16.232 4.0055
2.2274 0.017200 9.1884 0.97756 16.333 4.0576
2.3149 0.018878 9.2883 1.0041 16.433 4.1196
2.4037 0.021122 9.3890 1.0356 16.534 4.1757
2.4938 0.023054 9.4893 1.0617 16.635 4.2365
2.5843 0.025405 9.5898 1.0927 16.735 4.2967
2.6782 0.027826 9.6902 1.1235 16.836 4.3564
2.7706 0.030499 9.7908 1.1540 16.936 4.4154
2.8635 0.033377 9.8935 1.1841 17.135 4.5323
2.9572 0.036619 9.9941 1.2141 22.171 7.6974
3.0517 0.040245 10.095 1.2491 27.115 11.301
3.1464 0.043804 10.195 1.2839 32.150 14.834
3.2414 0.047587 10.296 1.3131 37.185 18.269
3.3370 0.051741 10.396 1.3474 42.228 21.756
3.4332 0.056436 10.496 1.3814 47.270 25.309
3.5298 0.061063 10.597 1.4151 52.310 29.000
3.6266 0.066121 10.697 1.4537 57.355 32.649
3.7233 0.071602 10.797 1.4920 62.394 36.422
3.8203 0.077023 10.897 1.5249 67.437 40.117
3.9177 0.083336 10.998 1.5627 72.478 43.696
4.0155 0.089187 11.098 1.5950 77.519 47.354
4.1131 0.096278 11.199 1.6322 82.560 50.983
4.2109 0.10351 11.299 1.6692 87.605 54.165
4.3114 0.11039 11.399 1.7110 92.645 57.312
4.4094 0.11826 11.499 1.7473 97.679 60.687
4.5078 0.12619 11.599 1.7834 102.72 64.238
4.6060 0.13486 11.700 1.8242 105.73 66.244
4.7047 0.14348 11.800 1.8649 115.76 73.770
4.8031 0.15253 11.901 1.9050 125.85 80.581
4.9019 0.16270 12.001 1.9450 135.90 86.421
5.0007 0.17256 12.101 1.9846 145.99 92.704
5.0996 0.18273 12.202 2.0290 156.08 98.812
5.1985 0.19347 12.302 2.0679 166.19 104.51
5.2978 0.20505 12.402 2.1117 176.30 110.05
5.3966 0.21646 12.503 2.1551 186.47 115.26
5.4956 0.22863 12.603 2.1930 196.59 120.25
5.5948 0.24118 12.704 2.2357 206.72 125.19
5.6945 0.25413 12.804 2.2832 216.82 131.33
5.7936 0.26769 12.904 2.3252 226.93 137.14
5.8931 0.28123 13.005 2.3720 237.02 142.90
5.9926 0.29621 13.105 2.4184 247.11 148.37
6.0921 0.31006 13.205 2.4645 257.17 153.60
6.1916 0.32517 13.306 2.5049 267.27 158.99
6.2915 0.33963 13.406 2.5554 277.36 165.66
6.3908 0.35666 13.507 2.6003 285.98 176.07
6.4906 0.37244 13.607 2.6447 286.21 179.95
6.5901 0.38877 13.707 2.6940 287.45 173.64
6.6898 0.40518 13.808 2.7376 291.13 176.92
6.7894 0.42601 13.908 2.7860 291.35 183.82
6.8891 0.44148 14.009 2.8339 296.08 177.07
6.9887 0.46201 14.109 2.8869 296.29 187.76
7.0887 0.48236 14.210 2.9339 297.56 177.53
7.1884 0.49736 14.310 2.9858 301.09 185.29
7.2885 0.51738 14.410 3.0373 301.30 189.24
7.3878 0.53729 14.511 3.0831 310.37 191.46
7.4878 0.55697 14.611 3.1336 310.38 192.50
7.5876 0.58171 14.712 3.1837 320.44 197.12
7.6876 0.60106 14.812 3.2385 320.45 199.20
7.7876 0.62543 14.913 3.2876 330.49 204.91
7.8876 0.64442 15.013 3.3414 330.50 206.57
7.9877 0.66842 15.127 3.4012 340.30 211.37
8.0875 0.69226 15.226 3.4544 340.53 216.02
8.1878 0.71588 15.327 3.5066 350.31 218.64
8.2872 0.73936 15.427 3.5584 350.32 224.28
8.3874 0.76780 15.528 3.6149 360.10 224.91
8.4874 0.78565 15.628 3.6656 360.33 229.90
8.5875 0.81370 15.729 3.7210 370.11 231.34
8.6873 0.84156 15.830 3.7759 370.12 239.35
8.7875 0.86919 15.930 3.8302
8.8876 0.89662 16.031 3.8892



image file: d4mr00124a-f2.tif
Fig. 2 Temperature dependence of the molar heat capacity CP of Y(BH4)3 in the temperature range from about 2 K to 370 K. The inset shows the relative deviation of the experimental data from the fit function.

The inset graph in Fig. 2 displays the relative deviation of the experimental data from the respective fit function. The relative deviation of the measured values from the fits is mostly within ±1.5%, but the deviations are rising to maximal ±4.5% below a temperature of about 5 K and above a temperature of about 280 K. In order to keep the degree of the fitting polynomials both as low as possible and to ensure still satisfying fits, the temperature range of our heat capacity measurements was divided into five appropriate intervals. Similarly as in our recently published works4–6,8,42,50–54 [Ca(AlH4)2, Sr(BH4)2], the heat capacities CP(T) were fitted then within these intervals using established polynomial functions.58 The applied functions and the corresponding temperatures are given below.

 
0 K to 5 K: CP = b·T + e·T3 + f·T5(4)
 
5 K to 17 K: CP = a + b·T + c·T2 + d·T−2 + e·T3 + g·T−1(5)
 
17 K to 100 K: CP = a + b·T + c·T2 + d·T−2 + e·T3(6)
 
100 K to 230 K: CP = a + b·T + c·T2 + d·T−2 + e·T3(7)
 
230 K to 380 K: CP = a + b·T + c·T2 + d·T2(8)

The results for the fit coefficients and the fit quality parameters R2 (coefficient of determination) and FitStdErr (fit standard error) are summarised in Table 2. The fits represent the experimental values with a deviation similar to the experimental standard uncertainty.

Table 2 Fitted coefficients of the heat capacity functions of Y(BH4)3 (M = 133.4341 g mol−1)
T interval [K] a [J mol−1 K−1] b [J mol−1 K−2] c [J K mol−1] d [J mol−1 K−3] e [J mol−1 K−4] f [J mol−1 K−6] g [J mol−1] R 2 FitStdErr
0 to 5 0 1.4814·10−3 0 0 1.2216·10−3 4.0961·10−6 0 1.0000 2.6436·10−4
5 to 17 6.0668·100 −8.0424·10−1 6.0859·10−2 3.8147·101 −8.6660·10−4 0 −2.4058·101 1.0000 3.1382·10−3
17 to 100 −6.5064·100 5.9847·10−1 2.2764·10−3 5.4218·101 −1.3670·10−5 0 0 0.9999 1.5525·10−1
100 to 230 −1.4167·102 2.7555·100 −1.1015·10−2 2.0107·105 1.8536·10−5 0 0 0.9999 2.6848·10−1
230 to 370 −2.2501·102 1.5227·100 −9.3269·10−4 3.3004·106 0 0 0 0.9879 3.0988·100


Applicability of Neumann–Kopp's rule to complex binding situations

Neumann–Kopp's rule is often used to calculate the heat unknown capacity of solid compounds from the heat capacities of the elements or smaller compounds forming the target compound.59–62 In the case of borohydrides, however, this method is not applicable because hydrogen is gaseous under standard conditions. Nevertheless, it should be possible to estimate their heat capacity function using Neumann–Kopp's rule by calculating the specific heat function of BH4 according to eqn (9) and using these values in combination with the specific heat of the metal of the studied metal borohydride to calculate its heat capacity according to eqn (10). The method was used for Mg(BH4)2 (ref. 61 and 63) and Ca(BH4)2.61 Dematteis et al.61 mentioned partly large deviations between the calculated and experimental values due to variations in the crystal structure and coordination number of the reactants. However, it remains unclear whether or not this method can be used for transition metal boranates, as there has been no application to this class of compounds so far.
 
CP,BH4 = CP,MBH4CP,M with M = Li, Na(9)
 
CP,M(BH4)x = CP,M + x·CP,BH4(10)

Table 3 lists the coefficients for the Maier Kelley58 heat capacity function polynomials of Y, Li and LiBH4 shown below:

 
CP = A + B·T + C·T−2 + D·T3 + E·T3 + F·T−1(11)

Table 3 Coefficients of the heat capacity functions according to eqn (11). The temperature ranges for the compounds were taken from the given sources
Compound T range [K] A [J mol−1 K] B [J mol−1 K−2] C [J K mol−1] D [J mol−1 K−3] E [J mol−1 K−4] F [J mol−1]
Y 20 to 300 (ref. 64) 2.831·10−1 4.842·10−2 1.098·104 −2.899·10−4 4.886·10−7 −1.093·103
300 to 1755 (ref. 65) 2.390·101 7.620·10−3 3.130·104 0 0 0
Li 20 to 105 (ref. 66) −1.615·100 2.728·10−2 2.814·102 2.085·10−3 −8.382·10−6 0
105 to 170 (ref. 66) 4.839·102 −7.341·100 −7.569·105 4.272·10−2 −8.641·10−5 0
170 to 300 (ref. 66) −4.353·101 5.752·10−1 2.167·105 −1.791·10−3 2.024·10−6 0
300 to 453 (ref. 65) 3.895·101 −7.098·10−2 −3.202·105 1.193·10−4 0 0
LiBH4 20 to 300 (ref. 67) −4.903·101 9.717·10−1 −6.683·103 −3.169·10−3 4.538·10−6 9.822·102
300 to 400 (ref. 65) −4.671·101 6.779·10−1 1.728·105 −8.409·10−4 0 0


The coefficients for the CP range below 300 K for Y and LiBH4 are derived from the fits of given literature values and the ones for Li are obtained from its G function. Above room temperature all values were taken from the HSC 5.1 database. All data used were validated by comparing various literature sources (as far as possible) and seem to posses high reliability. Respective references are shown in Table 3.

The comparison between our experimental values and those derived from the described method using LiBH4 are presented in Fig. 3. The use of NaBH4 as reported by Dematteis et al.61 was not reasonably applicable because of the presence of a phase transition at 189.9 K.68


image file: d4mr00124a-f3.tif
Fig. 3 Comparison between the heat capacity of Y(BH4)3 derived from our measurements (black solid line) and the one computed by applying Neumann–Kopp's rule according to ref. 61 and 63 (red dashed line: values below room temperature; blue dashed line: values above room temperature). The inset displays the relative deviation between the experimental values and the ones from the Neumann–Kopp rule.

The inset in Fig. 3 displays the deviation of the values of the experimental fit curve from the one calculated by the Neumann–Kopp rule. The deviation between both curves is significant and exceeds in a wide temperature range 10%. Towards high temperatures the slope of the Neumann–Kopp curve decreases so that both curves intersect at about 335 K. Similar behaviour has been found in other systems59,60,62 and was attributed to factors that are not reflected in the Neumann–Kopp rule such as magnetism and anharmonic effects60 or the formation of Schottky defects.62

In general Neumann–Kopp's rule is reasonably applicable for simple compounds (alloys), whose elements show in pure form the same crystal structure as the compounds to be described. In cases where the compound possesses complex subunits, it is necessary not only to consider the binding situation within the subunit but also that of the subunit in the target compound.59,61 In the case of the calculation of the heat capacity of Y(BH4)3 according to eqn (9) and (10), the requirements that the crystal structures of the components and the binding situation of the subunits should be the same is not fulfilled. Li crystallises in the body centred cubic W type structure (space group 229),69 Y in the hexagonal Mg type structure (space group 194),70 LiBH4 in a orthorhombic structure (space group 62)31 and Y(BH4)3 in a cubic structure (space group 205).31

In the literature, the influence of the thermal expansion of the compounds, among other things, was used for the extension of the Neumann–Kopp approximation in order to obtain more realistic CP values. Kumar et al. and Leitner et al. applied the extended Neumann–Kopp method to CaHCl and CaHBr59 and mixed oxides,60 respectively and obtained significant improvements. Unfortunately, this procedure is currently not applicable because of the missing thermal expansion coefficient for Y(BH4)3.

By the integration of the conventional Neumann–Kopp CP function divided by the temperature in the range from 5 K to 300 K, a value of the standard entropy of S° (300 K) = 183.8 J mol−1 K−1 was derived. This value differs about 9% from the experimental one determined in this study (vide infra). This result is a further indication of the limitation in regard to the use of the Neumann–Kopp approximation.

Based on the given explanations, the Neumann–Kopp rule can be seen as a rough estimation method and cannot replace reliable experimental data. From our findings for Y(BH4)3 its application appears also not recommended for precise thermodynamic calculations regarding other transition metal boranates that have not yet been prepared/investigated.

Calculation of the standard entropy, entropy of formation and of the enthalpy change

The calculation of the standard entropy values S° was carried out using the fitted polynomials for the heat capacity values in the temperature range of 2 K to 370 K using eqn (12).
 
image file: d4mr00124a-t3.tif(12)

The value for the standard entropy S° at 298.15 K was determined to be S° (298.15 K) = (168.9 ± 5.1) J mol−1 K−1, which is considerably higher than the value reported by Lee et al. for the room temperature α-phase of Y(BH4)3 (149.4 J mol−1 K−1) derived via DFT calculations.34 All calculated values of S° are presented in Table 4. In addition, the molar standard entropy of formation ΔFS° has been calculated from eqn (13) resulting in a value of ΔFS° (298.15 K) = −677.7 J mol−1 K−1. The values for the standard entropy of yttrium, boron and hydrogen were taken from the HSC database.65

 
image file: d4mr00124a-t4.tif(13)

Table 4 Molar thermodynamic functions of Y(BH4)3 (M = 133.4341 g mol−1) at selected temperatures between 5 K to 380 K and pressure of p = 100 kPa
T [K] C P [J mol−1 K−1] ΔS0T [J mol−1 K−1] ΔH0T/T [J mol−1 K−1] Φ [J mol−1 K] T [K] C P [J mol−1 K−1] ΔS0T [J mol−1 K−1] ΔH0T/T [J mol−1 K−1] Φ [J mol−1 K]
5 0.17291 0.06087 0.044012 0.016855 195 119.53 106.40 59.589 46.807
10 1.2194 0.44076 0.32539 0.11538 200 122.135 109.45 61.120 48.335
15 3.3375 1.3011 0.94792 0.35321 205 124.767 112.50 62.640 49.863
20 6.3997 2.6616 1.9117 0.74991 210 127.434 115.54 64.151 51.391
25 9.7512 4.4478 3.1428 1.3051 215 130.150 118.57 65.654 52.918
30 13.188 6.5283 4.5295 1.9988 220 132.927 121.60 67.151 54.444
35 16.687 8.8233 6.0157 2.8076 225 135.778 124.61 68.644 55.970
40 20.234 11.283 7.5708 3.7117 230 138.715 127.63 70.136 57.495
45 23.815 13.872 9.1765 4.6957 235 141.089 130.63 71.615 59.019
50 27.421 16.568 10.821 5.7474 240 144.024 133.64 73.093 60.543
55 31.039 19.351 12.494 6.8570 245 147.061 136.64 74.571 62.065
60 34.659 22.207 14.190 8.0167 250 150.189 139.64 76.052 63.586
65 38.271 25.124 15.904 9.2201 255 153.397 142.64 77.537 65.107
70 41.863 28.092 17.630 10.462 260 156.675 145.65 79.027 66.627
75 45.426 31.102 19.364 11.737 265 160.016 148.67 80.524 68.147
80 48.949 34.146 21.104 13.042 270 163.410 151.69 82.027 69.666
85 52.423 37.218 22.844 14.374 275 166.851 154.72 83.538 71.185
90 55.836 40.311 24.582 15.729 280 170.331 157.76 85.057 72.704
95 59.178 43.420 26.316 17.105 285 173.846 160.81 86.584 74.223
100 62.440 46.539 28.041 18.498 290 177.389 163.86 88.119 75.742
105 65.914 49.668 29.760 19.908 295 180.956 166.92 89.662 77.261
110 69.443 52.816 31.483 21.332 298.15 183.213 168.86 90.638 78.219
115 72.934 55.980 33.210 22.770 300 184.541 169.99 91.213 78.781
120 76.367 59.157 34.937 24.220 305 188.141 173.07 92.773 80.302
125 79.729 62.342 36.662 25.681 310 191.751 176.16 94.340 81.823
130 83.012 65.534 38.381 27.152 315 195.369 179.26 95.915 83.345
135 86.211 68.727 40.094 28.633 320 198.990 182.37 97.497 84.868
140 89.326 71.919 41.797 30.122 325 202.611 185.48 99.087 86.392
145 92.358 75.106 43.488 31.618 330 206.231 188.60 100.683 87.917
150 95.310 78.287 45.167 33.121 335 209.845 191.73 102.28 89.443
155 98.189 81.460 46.831 34.629 340 213.453 194.86 103.89 90.970
160 101.00 84.622 48.480 36.142 345 217.050 198.01 105.51 92.498
165 103.75 87.772 50.113 37.659 350 220.636 201.15 107.13 94.028
170 106.45 90.909 51.731 39.179 355 224.209 204.31 108.75 95.559
175 109.11 94.034 53.332 40.701 360 227.766 207.47 110.38 97.092
180 111.74 97.144 54.918 42.226 365 231.305 210.64 112.01 98.625
185 114.34 100.24 56.489 43.752 370 234.826 213.81 113.65 100.160
190 116.93 103.32 58.045 45.279


For the calculation of the change in molar enthalpy ΔH0T = H(T) − H(0) the heat capacity functions were integrated according to eqn (14).

 
image file: d4mr00124a-t5.tif(14)

The calculated values of ΔH0T divided by the temperature are given in Table 4 as well as the Φ parameter, which can be computed by eqn (15).

 
image file: d4mr00124a-t6.tif(15)
ΔS0T is equal to the absolute standard entropy S(T)° (see eqn (12)). However, it is necessary to carry out reliable calorimetric measurements to obtain the enthalpy of formation, which will then be used to construct the Gibbs energy function with the data obtained in this study.

Heat capacity at low temperatures

In the low temperature range between about 2 K and 5 K a three parameter fit function was applied (see eqn (16)). Fig. 4 displays the corresponding CP/T-vs.-T2 plot showing a high value of the coefficient of determination (R2) for the applied temperature range of the low temperature fit.
 
CP = CP,elec + CP,vib = γ·T + β·T3 + δ·T5(16)

image file: d4mr00124a-f4.tif
Fig. 4 Plot of CP/T versus T2 of Y(BH4)3 in the temperature range from 0 K to 5 K.

Usually, the linear term with the Sommerfeld coefficient γ is attributed to the electrical conductivity of metals and related compounds.50–54,71,72 However, there are also publications taking into account the electronic contribution to the heat capacity for non conductive materials. Loos et al.42 observed this behaviour for lithium iron phosphate and explained it with the existence of electronic states in structural defects. Defects in the crystal structure of the synthesised Y(BH4)3 that may potentially contribute to the heat capacity even at low temperatures can be assumed to be present because of the preparation by ball milling.71,73,74 Habermann et al. found that the linear term is also required for a sufficient fit of the heat capacity values at low temperatures for various metal alanates produced by ball milling.4,5,8 The need of a linear term for nonmetallic, insulating compounds in the low temperature range due to the existence of vacancies in the lattice was also postulated by Schliesser and Woodfield.74 The parameters β and δ in eqn (16) are related to the lattice vibration contribution to the heat capacity in terms of the Debye theory.71,72

A value of γ = (1.48 ± 0.01)·10−3 J mol−1 K−2 is obtained for the Sommerfeld coefficient, which is comparable to that one of metals.71 As mentioned above the phenomenon is common if structural defects are present in the sample.42,74 For the lattice vibration parameter, the value of β = (1.22 ± 0.01)·10−3 J mol−1 K−4 was determined. A Debye temperature of ΘD = (294.2 ± 1.1) K was calculated using eqn (17) with the given number of atoms per formula unit of n = 16 for Y(BH4)3. This value represents the excitation of low energy acoustic phonons (lattice vibrations), shown in the phonon density of states diagram calculated by Lee et al.34

 
image file: d4mr00124a-t7.tif(17)

The coefficient of the T5 term of the lattice contribution in eqn (16) was found to be δ = (4.09 ± 0.44)·10−6 J mol−1 K−6.

Conclusions

In the presented work, the heat capacity function of Y(BH4)3 between 2 K and 370 K was derived from mechanochemically prepared samples containing Y(BH4)3 + 3LiCl. The mixture was used as received after ball milling for the heat capacity measurements to avoid a partial decomposition of the boranate during the common solvent extraction. Our investigations show that this method is a convincing way for the determination of thermodynamic data from hydrides in mixtures obtained by mechanochemistry.

The measurement at low temperatures allows the determination of the absolute standard entropy resulting in a value of S° (298.15 K) = (168.9 ± 5.1) J mol−1 K−1 and a Debye temperature of ΘD = (294.2 ± 1.1) K. Despite the fact that the material is non-metallic, a comparably high value of γ = (1.48 ± 0.01)·10−3 J mol−1 K−2 was obtained for the Sommerfeld coefficient. This fact seems to be not unusual for mechanochemically prepared complex metal hydrides.

The absolute standard entropy derived from the measured data deviates significantly from the value of this quantity determined via DFT calculations in the literature. Given the determined uncertainty of the measurements of less than 5%, the obtained values appear to be more convincing. Although not surprising, however often applied, the estimate of the heat capacity via the modified Neumann–Kopp rule with the inclusion of complex subunits, in the presented case BH4, did not yield a satisfactory agreement with experimental findings.

Further investigations should address the determination of the enthalpy of formation as well as the understanding of the complex decomposition process. It is necessary to obtain thermodynamic values for possible decomposition intermediates to perform thermodynamic calculations on their existence during the decomposition process of Y(BH4)3. Applying this knowledge hopefully a deeper understanding of the potential use of Y(BH4)3 for reversible hydrogen storage concerning thermodynamic tuning strategies will be reached.

Data availability

The data supporting this article have been included as part of the manuscript. In detail, Table 1 gives all measured heat capacity values depending on the measurement temperature. The data to calculate the curves in Fig. 3 are calculated with the heat capacity values found in the literature, whereby all references indicated in Table 3.

Author contributions

K. Burkmann – investigation, formal analysis, validation, visualisation, writing – original draft. F. Habermann – visualisation, validation, writing – review & editing. B. Störr – investigation, writing – review & editing. J. Seidel – investigation, validation, writing – review & editing. K. Bohmhammel – formal analysis, supervision, validation, writing – review & editing. R. Gumeniuk – investigation, resources, writing – review & editing. F. Mertens – conceptualisation, funding acquisition, resources, supervision, project administration, writing – review & editing.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

The reported research activities have been financially supported by the Free State of Saxony (K. Burkmann, Landesstipendium zur Graduiertenförderung) and the Deutsche Forschungsgemeinschaft (DFG) within the project 449160425 (F. Habermann) and 422219901 (DynaCool-12 system).

Notes and references

  1. K. Suárez-Alcántara, J. R. Tena-Garcia and R. Guerrero-Ortiz, Materials, 2019, 12, 2724–2787 CrossRef PubMed.
  2. K. Suárez-Alcántara and J. R. T. García, Materials, 2021, 14, 2561 CrossRef PubMed.
  3. F. Habermann, K. Burkmann, B. Hansel, B. Störr, C. Schimpf, J. Seidel, M. Bertau and F. Mertens, Dalton Trans., 2023, 52, 4880–4890 RSC.
  4. F. Habermann, A. Wirth, K. Burkmann, B. Störr, J. Seidel, R. Gumeniuk, K. Bohmhammel and F. Mertens, ChemPhysChem, 2024, 25, e202300748 CrossRef CAS PubMed.
  5. F. Habermann, A. Wirth, K. Burkmann, J. Kraus, B. Störr, H. Stöcker, J. Seidel, J. Kortus, D. C. Meyer, R. Gumeniuk, K. Bohmhammel and F. Mertens, RSC Mechanochem., 2024 Search PubMed , submitted.
  6. K. Burkmann, F. Habermann, E. Schumann, J. Kraus, B. Störr, H. Schmidt, E. Brendler, J. Seidel, K. Bohmhammel, J. Kortus and F. Mertens, New J. Chem., 2024, 48, 2743–2754 RSC.
  7. Z. Cao, F. Habermann, K. Burkmann, M. Felderhoff and F. Mertens, Hydrogen, 2024, 5, 241–279 CrossRef CAS.
  8. F. Habermann, K. Burkmann, J. Kraus, B. Störr, J. Seidel, K. Bohmhammel, J. Kortus, R. Gumeniuk and F. Mertens, J. Alloys Compd., 2024, 980, 173476 CrossRef CAS.
  9. K. Burkmann, A. Demmer, F. Habermann, B. Hansel, B. Störr, J. Seidel, R. Gumeniuk, M. Bertau, K. Bohmhammel and F. Mertens, J. Therm. Anal. Calorim., 2025 DOI:10.1007/s10973-025-14085-z.
  10. C. Milanese, T. R. Jensen, B. C. Hauback, C. Pistidda, M. Dornheim, H. Yang, L. Lombardo, A. Zuettel, Y. Filinchuk, P. Ngene, P. E. de Jongh, C. E. Buckley, E. M. Dematteis and M. Baricco, Int. J. Hydrogen Energy, 2019, 44, 7860–7874 CrossRef CAS.
  11. E. M. Dematteis, M. B. Amdisen, T. Autrey, J. Barale, M. E. Bowden, C. E. Buckley, Y. W. Cho, S. Deledda, M. Dornheim, P. de Jongh, J. B. Grinderslev, G. Gizer, V. Gulino, B. C. Hauback, M. Heere, T. W. Heo, T. D. Humphries, T. R. Jensen, S. Y. Kang, Y.-S. Lee, H.-W. Li, S. Li, K. T. Møller, P. Ngene, S.-I. Orimo, M. Paskevicius, M. Polanski, S. Takagi, L. Wan, B. C. Wood, M. Hirscher and M. Baricco, Prog. Energy, 2022, 4, 032009 CrossRef CAS.
  12. R. Černý, F. Murgia and M. Brighi, J. Alloys Compd., 2022, 895, 162659 CrossRef.
  13. C. Comanescu, Materials, 2022, 15, 2286 CrossRef CAS PubMed.
  14. S. Di Cataldo and L. Boeri, Phys. Rev. B, 2023, 107, L060501 CrossRef CAS.
  15. Y. Yan, H.-W. Li, T. Sato, N. Umeda, K. Miwa, S.-i. Towata and S.-I. Orimo, Int. J. Hydrogen Energy, 2009, 34, 5732–5736 CrossRef CAS.
  16. J. J. Vajo, F. Mertens, C. C. Ahn, R. C. Bowman and B. Fultz, J. Phys. Chem. B, 2004, 108, 13977–13983 CrossRef CAS.
  17. J. J. Vajo, S. L. Skeith and F. Mertens, J. Phys. Chem. B, 2005, 109, 3719–3722 CrossRef CAS PubMed.
  18. T. Jaroń and W. Grochala, Dalton Trans., 2010, 39, 160–166 RSC.
  19. F. C. Gennari, Int. J. Hydrogen Energy, 2012, 37, 18895–18903 CrossRef CAS.
  20. T. Jaroń, W. Koźmiński and W. Grochala, Phys. Chem. Chem. Phys., 2011, 13, 8847–8851 RSC.
  21. T. Jaroń and W. Grochala, Dalton Trans., 2011, 40, 12808–12817 RSC.
  22. T. Jaroń, W. Wegner, K. J. Fijałkowski, P. J. Leszczyński and W. Grochala, Chem.–Eur. J., 2015, 21, 5689–5692 CrossRef PubMed.
  23. Y. Nakamori, H.-W. Li, M. Matsuo, K. Miwa, S. Towata and S. Orimo, J. Phys. Chem. Solids, 2008, 69, 2292–2296 CrossRef CAS.
  24. K. Park, H.-S. Lee, A. Remhof, Y.-S. Lee, Y. Yan, M.-Y. Kim, S. J. Kim, A. Züttel and Y. W. Cho, Int. J. Hydrogen Energy, 2013, 38, 9263–9270 CrossRef CAS.
  25. D. B. Ravnsbaek, Y. Filinchuk, R. Cerný, M. B. Ley, D. Haase, H. J. Jakobsen, J. Skibsted and T. R. Jensen, Inorg. Chem., 2010, 49, 3801–3809 CrossRef CAS PubMed.
  26. T. Sato, K. Miwa, Y. Nakamori, K. Ohoyama, H.-W. Li, T. Noritake, M. Aoki, S.-i. Towata and S.-I. Orimo, Phys. Rev. B:Condens. Matter Mater. Phys., 2008, 77, 104114 CrossRef.
  27. M. B. Ley, M. Paskevicius, P. Schouwink, B. Richter, D. A. Sheppard, C. E. Buckley and T. R. Jensen, Dalton Trans., 2014, 43, 13333–13342 RSC.
  28. M. Mostajeran, E. Ye, S. Desgreniers and R. T. Baker, Chem. Mater., 2017, 29, 742–751 CrossRef CAS.
  29. B. Richter, J. B. Grinderslev, K. T. Møller, M. Paskevicius and T. R. Jensen, Inorg. Chem., 2018, 57, 10768–10780 CrossRef CAS PubMed.
  30. A. Brukl and K. Rossmanith, Monatsh. Chem., 1959, 90, 481–487 CrossRef CAS.
  31. E. Roedern, Y.-S. Lee, M. B. Ley, K. Park, Y. W. Cho, J. Skibsted and T. R. Jensen, J. Mater. Chem. A, 2016, 4, 8793–8802 RSC.
  32. K. Rossmanith and E. Muckenhuber, Monatsh. Chem., 1961, 92, 600–604 CrossRef CAS.
  33. H. C. Brown, Y. M. Choi and S. Narasimhan, Inorg. Chem., 1982, 21, 3657–3661 CrossRef CAS.
  34. Y.-S. Lee, J.-H. Shim and Y. W. Cho, J. Phys. Chem. C, 2010, 114, 12833–12837 Search PubMed.
  35. C. Frommen, N. Aliouane, S. Deledda, J. E. Fonneløp, H. Grove, K. Lieutenant, I. Llamas-Jansa, S. Sartori, M. H. Sørby and B. C. Hauback, J. Alloys Compd., 2010, 496, 710–716 CrossRef CAS.
  36. R. Černý and P. Schouwink, Acta Crystallogr., Sect. B:Struct. Sci., Cryst. Eng. Mater., 2015, 71, 619–640 CrossRef PubMed.
  37. M. B. Ley, S. Boulineau, R. Janot, Y. Filinchuk and T. R. Jensen, J. Phys. Chem. C, 2012, 116, 21267–21276 CrossRef CAS.
  38. S. P. GharibDoust, M. Brighi, Y. Sadikin, D. B. Ravnsbæk, R. Černý, J. Skibsted and T. R. Jensen, J. Phys. Chem. C, 2017, 121, 19010–19021 CrossRef.
  39. C. A. Kennedy, M. Stancescu, R. A. Marriott and M. A. White, Cryogenics, 2007, 47, 107–112 CrossRef CAS.
  40. J. C. Lashley, M. F. Hundley, A. Migliori, J. L. Sarrao, P. G. Pagliuso, T. W. Darling, M. Jaime, J. C. Cooley, W. L. Hults, L. Morales, D. J. Thoma, J. L. Smith, J. Boerio-Goates, B. F. Woodfield, G. R. Stewart, R. A. Fisher and N. E. Phillips, Cryogenics, 2003, 43, 369–378 CrossRef CAS.
  41. S. Loos, Dissertation, TU Bergakademie Freiberg, Freiberg, 2014 Search PubMed.
  42. S. Loos, D. Gruner, M. Abdel-Hafiez, J. Seidel, R. Hüttl, A. U. Wolter, K. Bohmhammel and F. Mertens, J. Chem. Thermodyn., 2015, 85, 77–85 CrossRef CAS.
  43. Q. Shi, C. L. Snow, J. Boerio-Goates and B. F. Woodfield, J. Chem. Thermodyn., 2010, 42, 1107–1115 CrossRef CAS.
  44. C. A. Swenson, Rev. Sci. Instrum., 1999, 70, 2728–2731 CrossRef CAS.
  45. W. Schnelle, J. Engelhardt and E. Gmelin, Cryogenics, 1999, 39, 271–275 CrossRef CAS.
  46. G. K. White and S. J. Collocott, J. Phys. Chem. Ref. Data, 1984, 13, 1251–1257 CrossRef CAS.
  47. D. A. Shirley, J. Am. Chem. Soc., 1960, 82, 3841–3843 CrossRef CAS.
  48. D. F. Moyer, J. Phys. Chem. Solids, 1965, 26, 1459–1462 CrossRef CAS.
  49. W. G. Mallard and P. J. Linstrom, NIST Chemistry WebBook: NIST Standard Reference Database Number 69, Datenbank, 2022, https://webbook.nist.gov/ Search PubMed.
  50. D. Thomas, M. Abdel-Hafiez, T. Gruber, R. Hüttl, J. Seidel, A. U. Wolter, B. Büchner, J. Kortus and F. Mertens, J. Chem. Thermodyn., 2013, 64, 205–225 CrossRef CAS.
  51. D. Thomas, M. Zeilinger, D. Gruner, R. Hüttl, J. Seidel, A. U. Wolter, T. F. Fässler and F. Mertens, J. Chem. Thermodyn., 2015, 85, 178–190 CrossRef CAS.
  52. F. Taubert, S. Schwalbe, J. Seidel, R. Hüttl, T. Gruber, R. Janot, M. Bobnar, R. Gumeniuk, F. Mertens and J. Kortus, Int. J. Mater. Res., 2017, 108, 942–958 CAS.
  53. F. Taubert, J. Seidel, R. Hüttl, M. Bobnar, R. Gumeniuk and F. Mertens, J. Chem. Thermodyn., 2018, 116, 323–329 CrossRef CAS.
  54. F. Taubert, J. Seidel, R. Hüttl, M. Bobnar, R. Gumeniuk and F. Mertens, J. Chem. Thermodyn., 2019, 130, 119–128 CrossRef CAS.
  55. S. M. Sarge, G. W. H. Höhne and W. Hemminger, Calorimetry: Fundamentals, Instrumentation and Applications, Wiley-VCH Verlag, Weinheim, Germany, 2014 Search PubMed.
  56. G. Della Gatta, M. J. Richardson, S. M. Sarge and S. Stølen, Pure Appl. Chem., 2006, 78, 1455–1476 CrossRef CAS.
  57. G. Bergerhoff and I. D. Brown, Crystallographic Databases, International Union of Crystallography, Chester, 1987 Search PubMed.
  58. C. G. Maier and K. K. Kelley, J. Am. Chem. Soc., 1932, 54, 3243–3246 CrossRef CAS.
  59. S. S. Kumar, S. Ghosh, M. Sahu and R. Ganesan, J. Chem. Sci., 2019, 131, 1–7 CrossRef CAS.
  60. J. Leitner, P. Voňka, D. Sedmidubský and P. Svoboda, Thermochim. Acta, 2010, 497, 7–13 CrossRef CAS.
  61. E. M. Dematteis, S. R. Jensen, T. R. Jensen and M. Baricco, J. Chem. Thermodyn., 2020, 143, 106055 CrossRef CAS.
  62. T. Zienert and O. Fabrichnaya, Calphad, 2019, 65, 177–193 CrossRef CAS.
  63. E. R. Pinatel, E. Albanese, B. Civalleri and M. Baricco, J. Alloys Compd., 2015, 645, S64–S68 CrossRef CAS.
  64. G. A. Berezovskij, I. E. Paukov, G. S. Burkhanov and O. D. Chistyakov, Zh. Fiz. Khim., 1989, 63, 3141–3143 Search PubMed.
  65. A. Roine, HSC Chemistry, 2002, https://www.hsc-chemistry.com/ Search PubMed.
  66. P. Wang, A. Kozlov, D. Thomas, F. Mertens and R. Schmid-Fetzer, Intermetallics, 2013, 42, 137–145 CrossRef CAS.
  67. N. C. Hallett and H. L. Johnston, J. Am. Chem. Soc., 1953, 75, 1496–1497 Search PubMed.
  68. H. L. Johnston and N. C. Hallett, J. Am. Chem. Soc., 1953, 75, 1467–1468 CrossRef CAS.
  69. M. R. Nadler and C. P. Kempier, Anal. Chem., 1959, 31, 2109 CrossRef CAS.
  70. F. H. Spedding, A. H. Daane and K. W. Herrmann, Acta Crystallogr., 1956, 9, 559–563 CrossRef CAS.
  71. E. Gopal, Specific Heats at Low Temperatures, Springer, New York, NY, 1966 Search PubMed.
  72. A. Tari, The Specific Heat of Matter at Low Temperatures, Imperial College Press, London and River Edge, NJ, 2003 Search PubMed.
  73. R. Felten, U. Umhofer, U. Rauchschwalbe and G. Weber, J. Magn. Magn. Mater., 1985, 47–48, 39–41 CrossRef CAS.
  74. J. M. Schliesser and B. F. Woodfield, Phys. Rev. B:Condens. Matter Mater. Phys., 2015, 91, 024109 CrossRef.

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