The low-temperature heat capacity of fullerite C₆₀

The heat capacity at constant pressure of fullerite C₆₀ has been investigated using an adiabatic calorimeter in a temperature range from 1.2 to 120 K. Our results and literature data have been analyzed in a temperature interval from 0.2 to 300 K. The contributions of the intramolecular and lattice...

Повний опис

Збережено в:
Бібліографічні деталі
Дата:2015
Автори: Bagatskii, M.I., Sumarokov, V.V., Barabashko, M.S., Dolbin, A.V., Sundqvist, B.
Формат: Стаття
Мова:English
Опубліковано: Фізико-технічний інститут низьких температур ім. Б.І. Вєркіна НАН України 2015
Назва видання:Физика низких температур
Теми:
Онлайн доступ:http://dspace.nbuv.gov.ua/handle/123456789/127965
Теги: Додати тег
Немає тегів, Будьте першим, хто поставить тег для цього запису!
Назва журналу:Digital Library of Periodicals of National Academy of Sciences of Ukraine
Цитувати:The low-temperature heat capacity of fullerite C₆₀ / М.I. Bagatskii, V.V. Sumarokov, M.S. Barabashko, A.V. Dolbin, B. Sundqvist// Физика низких температур. — 2015. — Т. 41, № 8. — С. 812–819. — Бібліогр.: 54 назв. — англ.

Репозитарії

Digital Library of Periodicals of National Academy of Sciences of Ukraine
id irk-123456789-127965
record_format dspace
spelling irk-123456789-1279652018-01-01T03:03:35Z The low-temperature heat capacity of fullerite C₆₀ Bagatskii, M.I. Sumarokov, V.V. Barabashko, M.S. Dolbin, A.V. Sundqvist, B. Наноструктуры при низких температурах The heat capacity at constant pressure of fullerite C₆₀ has been investigated using an adiabatic calorimeter in a temperature range from 1.2 to 120 K. Our results and literature data have been analyzed in a temperature interval from 0.2 to 300 K. The contributions of the intramolecular and lattice vibrations into the heat capacity of C₆₀ have been separated. The contribution of the intramolecular vibration becomes significant above 50 K. Below 2.3 K the experimental temperature dependence of the heat capacity of C60 is described by the linear and cubic terms. The limiting Debye temperature at T → 0 K has been estimated (Θ0 = 84.4 K). In the interval from 1.2 to 30 K the experimental curve of the heat capacity of C₆₀ describes the contributions of rotational tunnel levels, translational vibrations (in the Debye model with Θ0 = 84.4 K), and librations (in the Einstein model with ΘE,lib = 32.5 K). It is shown that the experimental temperature dependences of heat capacity and thermal expansion are proportional in the region from 5 to 60 K. The contribution of the cooperative processes of orientational disordering becomes appreciable above 180 K. In the high-temperature phase the lattice heat capacity at constant volume is close to 4.5 R, which corresponds to the high-temperature limit of translational vibrations (3 R) and the near-free rotational motion of C60 molecules (1.5 R). 2015 Article The low-temperature heat capacity of fullerite C₆₀ / М.I. Bagatskii, V.V. Sumarokov, M.S. Barabashko, A.V. Dolbin, B. Sundqvist// Физика низких температур. — 2015. — Т. 41, № 8. — С. 812–819. — Бібліогр.: 54 назв. — англ. 0132-6414 PACS: 65.40.Ba, 65.80.–g, 81.05.ub http://dspace.nbuv.gov.ua/handle/123456789/127965 en Физика низких температур Фізико-технічний інститут низьких температур ім. Б.І. Вєркіна НАН України
institution Digital Library of Periodicals of National Academy of Sciences of Ukraine
collection DSpace DC
language English
topic Наноструктуры при низких температурах
Наноструктуры при низких температурах
spellingShingle Наноструктуры при низких температурах
Наноструктуры при низких температурах
Bagatskii, M.I.
Sumarokov, V.V.
Barabashko, M.S.
Dolbin, A.V.
Sundqvist, B.
The low-temperature heat capacity of fullerite C₆₀
Физика низких температур
description The heat capacity at constant pressure of fullerite C₆₀ has been investigated using an adiabatic calorimeter in a temperature range from 1.2 to 120 K. Our results and literature data have been analyzed in a temperature interval from 0.2 to 300 K. The contributions of the intramolecular and lattice vibrations into the heat capacity of C₆₀ have been separated. The contribution of the intramolecular vibration becomes significant above 50 K. Below 2.3 K the experimental temperature dependence of the heat capacity of C60 is described by the linear and cubic terms. The limiting Debye temperature at T → 0 K has been estimated (Θ0 = 84.4 K). In the interval from 1.2 to 30 K the experimental curve of the heat capacity of C₆₀ describes the contributions of rotational tunnel levels, translational vibrations (in the Debye model with Θ0 = 84.4 K), and librations (in the Einstein model with ΘE,lib = 32.5 K). It is shown that the experimental temperature dependences of heat capacity and thermal expansion are proportional in the region from 5 to 60 K. The contribution of the cooperative processes of orientational disordering becomes appreciable above 180 K. In the high-temperature phase the lattice heat capacity at constant volume is close to 4.5 R, which corresponds to the high-temperature limit of translational vibrations (3 R) and the near-free rotational motion of C60 molecules (1.5 R).
format Article
author Bagatskii, M.I.
Sumarokov, V.V.
Barabashko, M.S.
Dolbin, A.V.
Sundqvist, B.
author_facet Bagatskii, M.I.
Sumarokov, V.V.
Barabashko, M.S.
Dolbin, A.V.
Sundqvist, B.
author_sort Bagatskii, M.I.
title The low-temperature heat capacity of fullerite C₆₀
title_short The low-temperature heat capacity of fullerite C₆₀
title_full The low-temperature heat capacity of fullerite C₆₀
title_fullStr The low-temperature heat capacity of fullerite C₆₀
title_full_unstemmed The low-temperature heat capacity of fullerite C₆₀
title_sort low-temperature heat capacity of fullerite c₆₀
publisher Фізико-технічний інститут низьких температур ім. Б.І. Вєркіна НАН України
publishDate 2015
topic_facet Наноструктуры при низких температурах
url http://dspace.nbuv.gov.ua/handle/123456789/127965
citation_txt The low-temperature heat capacity of fullerite C₆₀ / М.I. Bagatskii, V.V. Sumarokov, M.S. Barabashko, A.V. Dolbin, B. Sundqvist// Физика низких температур. — 2015. — Т. 41, № 8. — С. 812–819. — Бібліогр.: 54 назв. — англ.
series Физика низких температур
work_keys_str_mv AT bagatskiimi thelowtemperatureheatcapacityoffulleritec60
AT sumarokovvv thelowtemperatureheatcapacityoffulleritec60
AT barabashkoms thelowtemperatureheatcapacityoffulleritec60
AT dolbinav thelowtemperatureheatcapacityoffulleritec60
AT sundqvistb thelowtemperatureheatcapacityoffulleritec60
AT bagatskiimi lowtemperatureheatcapacityoffulleritec60
AT sumarokovvv lowtemperatureheatcapacityoffulleritec60
AT barabashkoms lowtemperatureheatcapacityoffulleritec60
AT dolbinav lowtemperatureheatcapacityoffulleritec60
AT sundqvistb lowtemperatureheatcapacityoffulleritec60
first_indexed 2025-07-09T08:06:35Z
last_indexed 2025-07-09T08:06:35Z
_version_ 1837155902827790336
fulltext Low Temperature Physics/Fizika Nizkikh Temperatur, 2015, v. 41, No. 8, pp. 812–819 The low-temperature heat capacity of fullerite C60 М.I. Bagatskii, V.V. Sumarokov, M.S. Barabashko, and A.V. Dolbin B. Verkin Institute for Low Temperature Physics and Engineering of the National Academy of Sciences of Ukraine 47 Lenin Ave., Kharkov 61103, Ukraine E-mail: bagatskii@ilt.kharkov.ua B. Sundqvist Department of Physics, Umea University, SE-901 87 Umea, Sweden Received April 8, 2015, published online June 25, 2015 The heat capacity at constant pressure of fullerite C60 has been investigated using an adiabatic calorimeter in a temperature range from 1.2 to 120 K. Our results and literature data have been analyzed in a temperature inter- val from 0.2 to 300 K. The contributions of the intramolecular and lattice vibrations into the heat capacity of C60 have been separated. The contribution of the intramolecular vibration becomes significant above 50 K. Below 2.3 K the experimental temperature dependence of the heat capacity of C60 is described by the linear and cubic terms. The limiting Debye temperature at T → 0 K has been estimated (Θ0 = 84.4 K). In the interval from 1.2 to 30 K the experimental curve of the heat capacity of C60 describes the contributions of rotational tunnel levels, translational vibrations (in the Debye model with Θ0 = 84.4 K), and librations (in the Einstein model with ΘE,lib = 32.5 K). It is shown that the experimental temperature dependences of heat capacity and thermal expan- sion are proportional in the region from 5 to 60 K. The contribution of the cooperative processes of orientational disordering becomes appreciable above 180 K. In the high-temperature phase the lattice heat capacity at constant volume is close to 4.5 R, which corresponds to the high-temperature limit of translational vibrations (3 R) and the near-free rotational motion of C60 molecules (1.5 R). PACS: 65.40.Ba Heat capacity; 65.80.–g Thermal properties of small particles, nanocrystals, nanotubes and other related systems; 81.05.ub Fullerenes and related materials. Keywords: heat capacity, fullerite C60, lattice dynamics. Introduction Since the discovery of the fullerite molecule C60 [1], the low-temperature physical properties of fullerite C60 have been investigated by various methods: inelastic neu- tron scattering [2,3], infrared and Raman spectroscopy [4,5], x-ray [3,6,7], neutron [8] and electron [9,10] diffraction, NMR [11], dilatometry [12–15] and calorimetry [16–25]. It has found that fullerite C60 is a molecular crystal in which the molecules are bonded by the van der Waals forces and its physical properties are largely determined by the dynamics of the rotational motion of the C60 mole- cules. At Tc ≈ 260 K fullerite undergoes an orientational phase transition from a high-temperature face-centered cubic (FCC) lattice to a low-temperature simple cubic (SC) one [7,8]. The high-temperature phase has no long-range orientational order and the rotational motion of the mo- lecules is slightly hindered. In the low-temperature phase the centers of gravity of the molecules remain in FCC sites and molecules form four SC sublattices having different orientations of the axes of three-fold symmetry along the body diagonals of the cube (<111> directions). In the orien- tationally ordered phase by the rotation of the molecules about the <111> axes the molecules can be in six potential wells with global and local minima, which correspond to the pentagonal (p) and hexagonal (h) configurations, res- pectively. In the p-configuration one of the five-fold axes of the molecule is directed toward the middle of one of the double bonds of the neighboring molecule. In the h-confi- guration the three-fold axis of the molecule is directed to- ward the center of the double bond of the neighboring molecule [8,26]. The barrier between the wells is about © М.I. Bagatskii, V.V. Sumarokov, M.S. Barabashko, A.V. Dolbin, and B. Sundqvist, 2015 The low-temperature heat capacity of fullerite C60 2900 K and the energy difference between the minima in the potential wells of the p- and h-configurations is ≈ 120 K. Below Tc the C60 molecules either perform orientational vibrations (librations) in the potential wells or execute re- tarded rotation hopping between the nearest potential wells. The phase transition is cooperative in nature. The mo- lecule concentrations np and nh in the p- and h-configura- tions are dependent on temperature. In the low-temperature phase np ≈ 63% near Tc ≈ 260 K (see Fig. 10 in [2]). At lowering temperature np increase and the frequency of hopping decreases. At the temperature of glass formation Tg (80–90 K) the reorientational motion of the molecules is frozen and fullerite changes to the state of orientational glass with np ≈ 83% (see Fig. 10 [2]). This occurs because the energy of the molecules is no longer sufficient to over- come the potential barrier between the p- and h-configura- tions [2,8,27]. The presence of a high (≈ 2900 K) barrier between the p- and h-configurations places the emphasis on the influence of the temperature prehistory on the phys- ical properties of C60 crystals, which entails the hysteretic phenomena in the regions of phase transition. The heat capacity at constant pressure Cp(T) of C60 was investigated by the adiabatic method in Refs. 16, 17 (at a temperature range 11–300 K), [18] (13–300 K), [19] (5–340 K), [20] (6–350 K), [21] (1.2–30 K) and by thermal relaxation method in [22,23] (1.4–20 K), [24] (4–300 K) and [25] (0.2–190 K). In interval from 4 to 160 K the dis- crepancy between the data in [16–23] is within 25%. The data in [24] are 50–100% over those in [16–23]. Olson et al. [25] investigated a solid ~15% C70–C60 mixture. Therefore their Cp(T) differs significantly in value and behavior from the results in [16–24]. The systematic discrepancy between both the curves Cp(T) and the Tc — values taken on heat- ing the samples in adiabatic calorimeters [16–21] are due to the variations in the purity [16,17,28] and perfection [20] of the samples. The systematic scatter of the data on the thermodynamic properties of C60 is also caused by the influence of the temperature prehistory of the samples and by thermocycling [29–31]. Grivei et al. [24] observed a large hysteresis of Cp(T) in the region from 160 to 286 K. The heat capacity of C60 was investigated below 4 K [21–23,25]. Beyermann et al. [22,23] performed two series of Cp(T) measurements. In series 2 the samples were pre- annealed in vacuum at 430 K. Above 4 K the data discre- pancy between series 1 and 2 was within 16%. At T < 2 K the results of series 1 were an order of magnitude higher than in series 2. The distinctions between the Cp(T) data in series 1 [22] and 2 [23] were attributed [23] mainly to the different concentrations of the solvent impurity in the C60 samples. In the interval from 4 to 20 K the results of series 2 [23] are systematically 5–12% higher than the data in [21]. As temperature decreases, the data distinctions be- tween [21] and [23] increase and at 1.4 K the Cp in [23] is about five times higher than in [21]. Below 2.2 K [23] and 0.7 K [25] the data are described by the linear and cubic terms, respectively: 3 1 3( )pC T A T A T= + . (1) According to [23], the linear term in Cp(T) is more sen- sitive to the solvent impurity. It is determined by the con- tribution of the tunnel levels in the orientational glass phase of fullerite C60 [22,25]. The contribution described by the cubic term is fully dependent on the density of states of acoustic phonons. The limiting Debye temperature Θ0 = 80 K at T → 0 was estimated using A3 [25]. Beyer- mann et al. [23] and Nemes et al. [32] analyzed the expe- rimental results on Cp(T). By varying four parameters A1, Θ0, ΘE,tr and ΘE,lib they calculated the contributions of tunnel levels (A1), acoustic phonons within the Debye mo- del with the temperature Θ0 and optical translational and librations within the Einstein model with characteristic temperatures ΘE,tr and ΘE,lib, respectively. They obtained Θ0 = 37 K [23] and Θ0 = 32 K [32]. From the analysis of Cp(T) above 50 K Θ0 ≈ 50 K [16,17] and Θ0 ≈ 60 K [33]. Houen et al. [34] calculated Θ0 ≈ 100 K from the low- temperature Young modulus of single-crystalline C60. Θ0 ≈ 54 K was derived from the analysis of temperature dependence of the thermal expansion coefficient of poly- crystalline C60 and its solutions with Ne and Ar [14]. In [7] the ultrasonic velocities of polycrystalline C60 were ana- lyzed and extrapolated to low temperatures, which yielded Θ0 ≈ 55.4 K. Shebanovs et al. [35] obtained the Debye temperature ΘD(280 K) ≈ 53.9 K by analyzing the x-ray diffraction data for single-crystalline C60 at 280 K. ΘD(300 K) ≈ 66 K was obtained from the ultrasonic veloci- ties measured in a C60 single crystal at 300 K [36]. Mikhalchenko [37] analyzed literature data on Θ0 of fullerite C60. Author [37] calculated Θ0 = 77.12 K from the harmonic elastic constants cijkl of a C60 single crystal at 0 K. The large scatter of Θ0-data is a good motivation to continue low-temperature investigation of the physical pro- perties of fullerite C60. Θ0 is a constant which characterizes the properties of a crystal, such as heat capacity, electric and thermal conductivities, x-ray spectra intensity, elastic features. Θ0 is also a characteristic parameter separating the high temperature region (T >> Θ0), where the lattice vibrations can be described within the classical theory, and the low-temperature region (T << Θ0), where the quantum mechanical effects become significant [38]. Note that the data on the heat capacity of C60 are essen- tial for analyzing the heat capacities of new C60-containing nanomaterials, for example, elementary atomic and mole- cular gaseous substances that form interstitial solutions in the octahedral voids of fullerite C60 [39–41]. Goal of this study was to investigate the low-tempera- ture dynamics of fullerene C60 by the calorimetric method. Low Temperature Physics/Fizika Nizkikh Temperatur, 2015, v. 41, No. 8 813 М.I. Bagatskii, V.V. Sumarokov, M.S. Barabashko, A.V. Dolbin, and B. Sundqvist Experiment The heat capacity at constant pressure Cp(T) of fullerite C60 was investigated in a temperature interval from 1.2 to 120 K in an adiabatic calorimeter [21]. Two experiments were made. In the first experiment the heat capacity of the sample was measured from 1.2 to 30 K [21]. In the second experiment the sample was first heated to ~ 320 K and held in the dynamic vacuum (≈ 1·10–3 Torr) for about 48 hours. The calorimeter was cooled through wires without using helium as an heat exchange gas. Then measurements were made in the interval from 1.2 to 120 K. Cp(T) of fullerite C60 was obtained by subtracting addenda Cad(T) (the heat capacity of an empty calorimeter with the Apiezon grease) from the total heat capacity of the calorimeter with the sample. Cad(T) was measured in a special experiment. The sample was a cylinder about 6 mm high and 10 mm in diameter. It was prepared at Umea University (Sweden) by compacting a C60 powder under pressure about 1 kbar. The characteristic sizes of the C60 crystallites varied within 0.1 to 0.3 mm. The C60 purity was 99.99%. The masses of the C60 sample and the Apiezon grease were mf = (586.48 ± ± 0.05) mg and mA = (0.45 ± 0.05) mg, respectively. The in- formation about the C60 sample and the calorimeter is de- tailed elsewhere [14,21]. The contribution of the sample to the total heat capacity of the calorimeter with the sample was 45% below 2 K, about 70% in the interval from 4 to 20 K, 45% at 30 K and 23% at 120 K. The random experimental error in the spe- cific heat of fullerite C60 was ±40% at 1.3 K, ±30% at 2 K, ±5% at 4 K and ±1.5% in the interval from 30 to 120 K. Results and discussion The experimental results on the specific heat Cp(T) of fullerite C60 and the literature data [16–25] are illustrated in Fig. 1 for the temperature regions 0.2–300 K (a), 0.2–120 K (b) and 0.2–50 K (c) in the lg-lg scale representation. Our re- sults obtained in experiments 1 [21] and 2 coincide. Ac- cording to [7], the difference Cp(T)–Cv(T) is 0.18 mJ/(g·K) at 120 K, 1.17 mJ/(g·K) at 160 K and 4 mJ/(g·K) at 290 K, which makes about 0.1%, 0.4% and 0.6% of Cp(T), respec- tively. The difference Cp(T)–Cv(T) is negligible below 160 K. It is seen that two experimental curves [24,25] are sig- nificantly distinct from the others. There is a giant hystere- sis in temperature range from 160 to 286 K [24]. The curve Cp(T) measured on heating the calorimeter has a maximum at 286 K and a minimum at 255 K. The curve Cp(T) taken on cooling the calorimeter has two maxima at 157 and 252 K and two minima at 202 and 266 K [24] (see Fig. 1(a)). For our opinion the distinctions between [25] and [16–24] is the effect of the impurity ~15% C70 in the sample. Note that above 40 K the contribution of the impurity C70 is higher even at low concentrations because the C70 mole- cule has 30 intramolecular frequencies more than the C60 molecule. Fig. 1. (Color online) The temperature dependence of the specific heat of C60 in temperature range 0.2–300 K (a), 0.2–120 K (b), and 0.2–50 K (c), lg-lg scale. Experimental results: (○) — this work and [21], (+) — [16,17], (∆) — [18], (×) — [19], () — [20], (∇) — [23], (■) — [24], (●) — [25]. The contribution of intra- molecular vibrations Cin (solid line) to the heat capacity of C60 was calculated within the Einstein model using the vibration fre- quencies of the carbon atoms in the C60 molecule [42]. 814 Low Temperature Physics/Fizika Nizkikh Temperatur, 2015, v. 41, No. 8 The low-temperature heat capacity of fullerite C60 It is seen (Figs. 1(a),(b)) that our results agree well with the data of Atake et al. [16,17] (in interval 11–120 K) and Beyermann et al. [23] (4–20 K). At T < 4 K the distinction between our results and [23] increases. Near 1.5 K the data of [23] are five times higher than our results. Near 1.4 K our Cp values are close to those in [25] (see Fig. 1(c)). In the interval from 8 to 40 K the data of [18–20] are on the average 8% higher than our results. Note that in [16–20] the samples were hermetically sealed in calorimeter ves- sels, and the sample-calorimeter thermal contact was im- proved by feeding helium gas to the calorimeter at room temperatures. At this temperature the He atoms can occupy the octahedral voids in the C60 crystal [43,44]. Saturation of C60 with He takes several hours [44]. The presence of helium in the calorimetric vessels and the samples can lead to higher systematic errors. In our experiments the calori- metric cell was cooled through wires, without using helium as an heat exchange gas. Above 40 K the distinctions in the data of [16–25] increase. This may be caused by the presence of the C70 impurity (this work, Fig. 1(a) and [28]) and the solvent (see Fig. 2 in [17]) or disturbance of the equilibrium concentrations np and nh at Tg as well as by helium gas. The heat capacity of a C60 single crystal with a mini- mum of structural defects was measured in a temperature range from 6 to 350 K by Miyazaki et al. [20]. They ob- served the sharpest maximum in the curve Cp(T) and high- est phase transition temperature Tc = 262.1 K. It was found that the transition from the orientational glass state to a par- tially orientationally-ordered phase occurred in a range from 80 to 90 K. The difference between the heat capaci- ties above and below Tg ≈ 84.6 K is ΔC = 3.6 J/(K mol). The corresponding data in [18] and [19] are ΔC = = 7 J/(K·mol) at Tg ≈ 86.8 K and ΔC = 4.5 J/(K·mol) at Tg ≈ 86 K, respectively. Below Tg the C60 concentrations in the p- and h-configurations are frozen (np ≈ 83% [2]). Above Tg the concentration np(T) decreases when the tem- perature rises (see Fig. 10 in [2]). The decrease/increase in np is attended with heat absorption/release in the crystal [18,19]. The data on heat capacity [18] and thermal con- ductivity [45] show that the characteristic time τ of p–h relaxation increases from ~103 s at 90 K to ~104 s at 80 K. Our calorimeter was cooled from 150 to 80 K rather fast (2.5 hours). Therefore the concentration np frozen in the sample below 80 K can correspond to the equilibrium con- centration at Tf which is higher than Tg. As a result, the heat capacities measured at T < Tf can be lower than in the case of the equilibrium concentration np(Tf). In the region of the glass phase transition the scatter of data is several times higher than at T < 70 K and T > 90 K (see Fig. 1(b)). This may be attributed to the influence of the temperature prehistory of the sample. The results obtained and the literature data in tempera- ture range from 0.2 to 300 K [16–25] were analyzed as- suming an additive contribution of translational, rotational and intramolecular degrees of freedom to the heat capacity of fullerite. Accoding to the group-theory analysis [46], 174 intramolecular vibrations may be grouped into 46 fun- damental modes having characteristic symmetries: Ag (two modes), Au (one mode), T1g (three modes), T1u (four modes), T2u (five modes), Gg (six modes), Hg (eight modes), Hu (seven modes). The contribution of the intramolecular vib- rations Cin(T) was calculated using the Einstein model and the data on the vibration frequencies of the carbon atoms in a C60 molecule [42]. Cin(T) is illustrated in Fig. 1 (solid line). The contribution Cin(T) becomes appreciable above 50 K. Cin/Cp ≈ 0.5 at ≈ 100 K (the low-temperature phase) and Cin/Cp > 0.9 at 270 K (the high-temperature phase) (see Fig. 1(a)). The lattice heat capacity Cp,lat(T) = Cp(T) – – Cin(T) of fullerite C60 is illustrated in Fig. 2. It is seen that Cp,lat is weakly dependent on temperature in the inter- val from 60 to 150 K. The contribution of the cooperative processes of orientational disordering to the lattice heat ca- pacity of C60 is appreciable above 180 K. The peak in the curve Cp,lat(T) at Tc ≈ 260 K is due to the orientational order-disorder phase transition. At 290 K the CV,lat value is close to 4.5 R (R is a gas constant). CV,lat was obtained from the Cp,lat data [20] by subtracting the correction Cp–CV ≈ ≈ 3.0 J/(K·mol) [7]. This behavior of the heat capacity is consistent with the data in [2,48] which suggest that the rotation of C60 molecules is nearly free in the high-tem- perature phase. At low temperatures the experimental Cp,lat(T) was ana- lyzed taking into account the contributions of the rotational tunnel levels in orientational glass (Ctun), translational (CD) and libration (CE,lib) vibrational modes. Below 2.3 K our temperature dependence Cp,lat(T) is described by the expres- sion 3 ,lat ( ) 0.01 0.00322pC T T T= + (J/(K·mol)) (see Fig. 3). The linear term describes the contribution of the rotational tunnel levels in orientational C60 glass [15,22,23,25,48,49]. This term is sensitive to even low impurity concentrations [23,25]. The impurities randomly distributed in the crystal generate random deformation fields leading to higher stochastization of tunnel levels [50,51]. Fig. 2. (Color online) The lattice specific heat Cp,lat(T) = = Cp(T) – Cin(T) of fullerite C60. Experimental results: (○) — this work and [21], (+) — [16,17], (∆) — [18], (×) — [19], () — [20], (∇) — [23]. Low Temperature Physics/Fizika Nizkikh Temperatur, 2015, v. 41, No. 8 815 М.I. Bagatskii, V.V. Sumarokov, M.S. Barabashko, A.V. Dolbin, and B. Sundqvist The contribution described by the cubic term 3(0.00322T J/(K·mol) at T ≤ 2.3 K) is made by acoustic translational vibrations. The limiting Debye temperature at T → 0 K (Θ0 = 84.4 K) was found using the coefficient at T3. This value agrees with literature Θ0 = 80 K obtained for solid solution 15% C70 in C60 in experiment [25]. Cal- culation of Debye temperature Θ0 = 77.12 K was made in [37] by using the harmonic elastic constants cijkl of a C60 single crystal at 0 K. Literature data for Debye tem- perature are given in the Table 1. This scatter (unusual for solids) in the quantities Θ0 for the fullerite C60 is most likely associated not only with the limitations of the Debye law T3 (for the C60 the corresponding range is extremely narrow, T ≤ 2 K) but also with the procedures used for cal- culation. The contribution CD of translational vibration modes above 2.3 K was calculated within the Debye model with Θ0 = 84.4 K. The contribution of libration modes becomes significant above 2.3 K. The contribution of optical libra- tion modes CE,lib was calculated within the Einstein model. The best fitting between the experimental Cp,lat and calcu- lated ,lat 1 0 ,lib ,lib( ) ( , ) ( , )V D E EC T A T C T C T= + Θ + Θ curves in temperature interval from 1.2 to 30 K is observed for A1 = 0.01 J/K2mol, Θ0 = 84.4 K and ΘE,lib = 32.5 K. The lattice heat capacities Cp,lat and CV,lat of C60 are com- pared in Fig. 4. It is seen that the contributions of the acoustic translational modes and the tunnel levels become dominant below 3 K. The contribution of the librations prevails in the interval from 4 to 20 K. The densities of states of lattice vibrations (arbitrary units) are illustrated schematically in Fig. 5. The density of states (solid line) was obtained in experiment on inelastic neutron scattering [2]. The dotted and solid lines show the densities of states of the translational modes (Debye model) and optical libration modes (Einstein model) obtained from the analysis of our Cp,lat(T) data. It is seen in Fig. 5 that the first libration (L) maximum in the density of states curve is close to the Einstein temperature ΘE,lib = 32.5 K. The behavior of the temperature dependences of the heat capacity Cp(T) and the linear thermal expansion α(T) [14] Table 1. Debye temperature Authors Year Θ0, К Property Temperature range, K T. Atake et al. 1992 50* Heat capacity 10–300 W.P. Beyermann et al. [23] 1992 37 Heat capacity 1.4–20 J.R. Olson et al. [25] 1993 80 Heat capacity 0.2–190 N. A. Aksenova et al. [7] 1999 55.2* Sound velocity 300 N. P. Kobelev et al. [36] 1998 66 Sound velocity 300 A.N. Aleksandrovskii et al. [14] 2003 54** Heat expansion 6–12 S. Hoen et al. [34] 1992 100 Young’s modulus 80–120 M. N. Magomedov et al. 2005 58.75 Calculation 0 V. P. Mikhal’chenko [37] 2010 77.12 Calculation from the harmonic elastic constants 0 This paper 2015 84.4 Heat capacity 1–120 C o m m e n t s : * — From analysis of high-temperature data ** — From analysis of temperature dependencies of heat expansion of C60 and C60 with Ne and Ar admixtures. Fig. 4. (Color online) The lattice specific heat Cp,lat(T) = = Cp(T) – Cin(T) of C60. Experiment: (1) — this work and [21]. Calculation: (3) dot line — contribution of tunnel levels and trans- lational vibrations of the lattice (A1T+CD), (4) dash line — contri- bution of librations (CE,lib) and (2) solid line — the total heat ca- pacity. ,lat 1 0 ,lib ,lib( ) ( , ) ( , )V D E EC T A T C T C T= + Θ + Θ . Fig. 3. (Color online) The low-temperature heat capacity of C60 in the coordinates C/T vs T2. 816 Low Temperature Physics/Fizika Nizkikh Temperatur, 2015, v. 41, No. 8 The low-temperature heat capacity of fullerite C60 of C60 are compared in Fig. 6 in a range from 1 to 20 K. The linear thermal expansion was scaled along the ordinate axis to match the curves Cp(T) and α(T). The measure- ments of Cp(T) and α(T) were made on the same C60 sample. Two series of α(T) measurement were made [14]. The re- sults are illustrated in Fig. 6 (first — dashed line, second — solid line). The second series was made after series 1 and annealing the sample in vacuum (1 mTorr) at 450°C for 72 hours. It is seen that the temperature behavior of Cp(T) and α(T) measured on the non-annealed and annealed samples is similar in interval from 5 to 20 K. In this tem- perature region the contribution of the orientational molec- ular vibrations (librations) to Cp,lat(T) is dominant and therefore α(T) is mainly determined by the anharmonicity of the orientational vibrations. Below 5 K the curves Cp(T) and α(T) are close only for the annealed sample. The val- ues of α(T) of the no annealed sample are negative at liquid helium temperatures. We attribute this to the influence of the impurities and defects present in the sample. The linear thermal expansion α(T) is very sensitive to these factors in C60 below 5 K [14]. The data on α(T) of a C60 single crystal [52] and Cp(T) of our C60 powder sample are compare in the region from 5 to 120 K in Fig. 7. α(T) and Cp(T) exhibit similar tem- perature dependences in the interval from 5 to 63 K. Above 60 K the curve Cp(T) goes upwards due to the increasing con- tribution of the intramolecular vibrations (see Figs. 1(a), (b) and Fig. 2). The ratio 3α/Cp determines the temperature dependence of the thermodynamic Grüneisen parameter 3( ) V T VT C α γ = χ , (2) because the relation between the molar volume V(T) and the isothermal compressibility χT(T) is only slightly tem- perature dependent at low temperatures. The correction Cp–CV is negligible at 60 K (0.2 J/(K·mol) [7]). As seen in Figs. 6 and 7, γ(T) of fullerite C60 is only slightly depend- ent on temperature in the interval from 5 to 60 K. Accord- ing to [53] γ(T) ≈ 3. The behavior of α(T) and γ(T) below 5 K is discussed in [14,48,49]. In the interval from 70 to 100 K the curve α(T) taken on a C60 single crystal [52] has a feature – a sharp rise and a jump of α(T) at Tg ≈ 86 K (see Fig. 7). This feature of α(T) is also observed in x-ray and neutron diffraction studies (see Fig. 1 in [7]). Figure 7 suggests that the ratio α/Cp and hence γ(T) have a jump in the region of the glass phase transition. Jumps of γ(T) during orientational phase transi- tions (Tg ≈ 86 K, Tc ≈ 260 K) were observed in photoaco- ustic experiments [54]. The anomalous values of γ(T) (and other physical properties) are essentially lower during the glass phase transition than in the case of an order-disorder phase transition. Fig. 5. (Color online) The density of states of the lattice vibra- tions measured on the C60 powder by the method of inelastic neutron scattering (solid curve). L and T refer to libration and translation features, respectively The densities of state derived from the analysis of our Cp,lat(T) data are shown for the acoustic mode in the Debye model (straight dashed line), the optical libra- tion mode in the Einstein model (straight solid line). Fig. 6. (Color online) The comparison of the experimental curves Cp(T) ((○) — our data) and the linear thermal expansion α(T) (dashed line — series 1, solid line — series 2) [14] of fulle- rite C60. Fig. 7. (Color online) The comparison of the experimental curves Cp(T) ((○) — our data) and α(T) (solid line — [52]) of fullerite C60. Low Temperature Physics/Fizika Nizkikh Temperatur, 2015, v. 41, No. 8 817 М.I. Bagatskii, V.V. Sumarokov, M.S. Barabashko, A.V. Dolbin, and B. Sundqvist Conclusion The heat capacity of fullerite C60 at constant pressure has been investigated in the interval from 1.2 to 120 K using an adiabatic calorimeter. Our results and literature data were analyzed in the temperature interval from 0.2 to 300 K. The contributions of intramolecular and lattice vibrations to the heat capacity C60 have been separated. The contribution of intramolecular vibrations becomes essential above 50 K, being 50% at 100 K and over 90% above 270 K. The contribution of acoustic modes to the heat capacity of C60 is dominant below 2.3 K. The Debye temperature at T → 0 K has estimated (Θ0 = 84.4 K). The scatter of val- ues of Θ0 in literature data is caused mainly due from cal- culations involving the physical properties of C60 above 4 K as well as from the analysis procedures. In the interval from 1.2 to 40 K the experimental curve of the heat capacity of the C60 lattice is caused the contri- butions of the rotational tunnel levels, translational vibra- tions (within the Debye model with Θ0 = 84.4 K) and lib- rations (within the Einstein model with ΘE,lib = 32.5 K). ΘE,lib agrees well with the libration maximum in the curve of the density of states estimated by the method of inelastic neutron scattering. It is found that the experimental temperature depend- ences of heat capacity and thermal expansion are propor- tional in the region from 5 to 60 K. This suggests that the Grüneisen coefficient γ(T) of fullerite C60 is only slightly dependent on temperature in this interval. The features of the temperature dependences of the heat capacity of C60 in the regions of the heat capacity of C60 in the regions of the orientational glass Tg and order-disorder Tc phase transitions are determined by the potential field structure and the presence of p- and h-molecule configura- tions. In the orientationally ordered phase by the rotation of the molecules about the <111> axes the molecules can be in six potential wells with global and local minima, which cor- respond to the pentagonal (p) and hexagonal (h) configura- tions, respectively. The presence of a high (≈ 2900 K) bar- rier between the p- and h-configurations places the emphasis on the influence of the temperature prehistory on the phys- ical properties of C60 crystals, which entails the hysteretic phenomena in the regions of phase transition. In the high-temperature phase the specific heat of the lat- tice at constant volume is close to 4.5 R, which corre- sponds to the high-temperature limit for the translational lattice vibrations (3 R) and the near-free rotational motion of C60 molecules (1.5 R). Acknowledgments The authors are indebted to A.I. Prokhvatilov, A.I. Kriv- chikov and S.B. Feodosyev for a fruitful discussion. 1. H.W. Kroto, J.R. Heath, S.C. O'Brien, R.F. Curl, and R.E. Smalley, Nature 318, 162 (1985). 2. S. Rols, C. Bousige, J. Cambedouzou, P. Launois, J.-L. Sauvajol, H. Schober, V.N. Agafonov, V.A. Davydov, and J. Ollivier, Eur. Phys. J. Special Topics 213, 77 (2012). 3. P.A. Heiney, G.B.M. Vaughan, J.E. Fisher, N. Coustel, D.E. Cox, J.R.D. Copley, D.A. Neumann, W.A. Kamitakahara, K.M. Creegan, D.M. Cox, J.P. McCauley, Jr., and A.B. Smith III, Phys. Rev. B 45, 4544 (1992). 4. S.G. Stepanian, V.A. Karachevtsev, A.M. Plokhotnichenko, L. Adamowicz, and A.M. Rao, J. Phys. Chem. B 110, 15769 (2006). 5. A.V. Peschanskii, A.Yu Glamazda, V.I. Fomin, and V.A. Karachevtsev, Fiz. Nizk. Temp. 38, 1077 (2012) [Low. Temp. Phys. 38, 854 (2012)]. 6. L.S. Fomenko, V.D. Natsik, S.V. Lubenets, V.G. Lirtman, N.A. Aksenova, A.P. Isakina, A.I. Prokhvatilov, M.A. Strzhemechny, and R.S. Ruoff, Fiz. Nizk. Temp. 21, 465 (1995) [Low Temp. Phys. 21, 364 (1995)]. 7. N.A. Aksenova, A.P. Isakina, A.I. Prokhvatilov, and M.A. Strzhemechny, Fiz. Nizk. Temp. 25, 964 (1999) [Low Temp. Phys. 25, 724 (1999)]. 8. W.I.F. David, R.M. Ibberson, T.J.S. Dennis, J.P. Hare, and K. Prasside, Europhys. Lett. 18, 219 (1992). 9. A.T. Pugachev, N.P. Churakova, N.I. Gorbenko, Kh. Saadli, and E.S. Syrkin, Zh. Eksp. Teor. Fiz. 114, 1014 (1998). 10. A.T. Pugachev, N.P. Churakova, and N.I. Gorbenko, Fiz. Nizk. Temp. 23, 854 (1997) [Low Temp. Phys. 23, 642 (1997)]. 11. R.D. Johnson, C.S. Yannoni, H.C. Dorn, J.R. Salem, and D.S. Bethune, Science 255, 1235 (1992). 12. A.N. Aleksandrovskii, V.B. Esel'son, V.G. Manzhelii, B.G. Udovichenko, A.V. Soldatov, and B. Sundqvist, Fiz. Nizk. Temp. 23, 1256 (1997) [Low Temp. Phys. 23, 943 (1997)]. 13. A.N. Aleksandrovskii, V.B. Esel'son, V.G. Manzhelii, B.G. Udovichenko, A.V. Soldatov, and B. Sundqvist, Fiz. Nizk. Temp. 26, 100 (2000) [Low Temp. Phys. 26, 75 (2000)]. 14. A.N. Aleksandrovskii, A.S. Bakai, A.V. Dolbin, V.B. Esel’son, G.E. Gadd, V.G. Gavrilko, V.G. Manzhelii, S. Moricca, B. Sundqvist, and B.G. Udovidchenko, Fiz. Nizk. Temp. 29, 432 (2003) [Low Temp. Phys. 29, 324 (2003)]. 15. A.N. Aleksandrovskii, A.S. Bakai, A.V. Dolbin, D. Cassidy, V.G. Manzhelii, and B. Sundqvist, Fiz. Nizk. Temp. 31, 565 (2005) [Low Temp. Phys. 31, 429 (2005)]. 16. T. Atake, T. Tanaka, H. Kawaji, K. Kikuchi, K. Saito, S. Suzuki, I. Ikemoto, and Y. Ashiba, Physica C 185–189, 427 (1991). 17. T. Atake, T. Tanaka, H. Kawaji, K. Kikuchi, K. Saito, S. Suzuki, Y. Ashiba, and I. Ikemoto, Chem. Phys. Lett. 196, 321 (1992). 18. T. Matsuo, H. Suga, W.I.F. David, R.M. Ibberson, P. Bernier, A. Zahab, C. Farbe, A. Rassat, and A. Dworkin, Solid State Commun. 83, 711 (1992). 19. B.V. Lebedev, K.B. Zhogova, T.A. Bykova, B.S. Kaverin, V.L. Karnatsevich, and M.A. Lopatin, Rus. Chem. Bull. 45, 2113 (1996). 818 Low Temperature Physics/Fizika Nizkikh Temperatur, 2015, v. 41, No. 8 The low-temperature heat capacity of fullerite C60 20. Y. Miyazaki, M. Sorai, R. Lin, A. Dworkin, H. Szwarc, and J. Godard, Chem. Phys. Lett. 305, 293 (1999). 21. M.I. Bagatskii, V.V. Sumarokov, and A.V. Dolbin, Fiz. Nizk. Temp. 37, 535 (2011) [ Low Temp. Phys. 37, 424 (2011)]. 22. W.P. Beyermann, M.F. Hundley, J.D. Thompson, F.N. Diederich, and G. Grüner, Phys. Rev. Lett. 68, 2046 (1992). 23. W.P. Beyermann, M.F. Hundley, J.D. Thompson, F.N. Diederich, and G. Grüner, Phys. Rev. Lett. 69, 2737 (1992). 24. E. Grivei, M. Cassart, J.-P. Issi, L. Langer, B. Nysten, J.-P. Michenaud, C. Farbe, and A. Rassat, Phys. Rev. B. 44, 8514 (1993). 25. J.R. Olson, K.A. Topp, and R.O. Pohl, Science 259, 1145 (1993). 26. Science of Fullerenes and Carbon Nanotubes, M.S. Dresselhaus, G. Dresselhaus, and P.C. Eklund (eds.), Academic Press, San Diego (1996). 27. V.D. Natsik and A.V. Podolskiy, Fiz. Nizk. Temp. 26, 304 (2000) [Low Temp. Phys. 26, 225 (2000)]. 28. E. Grivei, B. Nysten, M. Cassart, J-P. Issi, C Farbe, and A. Rassat, Phys. Rev. B 47, 1705 (1993). 29. V.B. Efimov, L.P. Mezhov-Deglin, and R.K. Nikolaev, Pis’ma Zh. Eksp. Teor. Fiz. 65, 651 (1997) [JETP 65, 687 (1997)]. 30. G. Pitsi, J. Caerels, and J. Thoen, Phys. Rev. B 55, 915 (1997). 31. J.E. Fisher, A.R. McGhie, J.K. Estrada, M. Haluška, H. Kuzmany, and H.-U. ter Meer, Phys. Rev. B 53, 11418 (1996). 32. N.M. Nemes, M. García-Hernández, G. Bortel, É. Kováts, B.J. Nagy, I. Jalsovszky, and S. Pekker, J. Phys. Chem. B 113, 2042 (2009). 33. B. Sundqvist, Physica B 265, 208 (1999). 34. S. Hoen, N.G. Chopra, X.-D. Xiang, R. Mostovoy, J. Hou, W.A. Vareka, and A. Zettl, Phys. Rev.B 46, 12737 (1992). 35. L. Shebanovs, J. Maniks, and J. Kalnacs, J. Cryst. Grow. 234, 202 (2002). 36. N.P. Kobelev, R.K. Nikolaev, Ya.M. Soĭfer, and S.S. Khasanov, Fiz. Tverd. Tela 40, 173 (1998) [Phys. Solid State 40, 154 (1998)]. 37. V.P. Mikhal’chenko, Fiz. Tverd. Tela 52, 1444 (2010) [Phys. Solid State 52, 1549 (2010)]. 38. Theory of Anharmonic Effects in Crystals, G. Leibfried and W. Ludwig (eds.), Academic Press Inc., New York-London (1961). 39. M.I. Bagatskii, V.V. Sumarokov, A.V. Dolbin, and S. Sundqvist, Fiz. Nizk. Temp. 38, 87 (2012) [Low Temp. Phys. 38, 67 (2012)]. 40. M.I. Bagatskii, V.G. Manzhelii, V.V. Sumarokov, A.V. Dolbin, M.S. Barabashko, and B. Sundqvist, Fiz. Nizk. Temp. 40, 873 (2014) [Low Temp. Phys. 40, 678 (2014)]. 41. A.M. Gurevich, A.V. Terekhov, D.S. Kondrashev, A.V. Dolbin, D. Cassidy, G.E. Gadd, S. Moricca, and B. Sundqvist, Fiz. Nizk. Temp. 32, 1275 (2006) [Low Temp. Phys. 32, 967 (2006)]. 42. V. Schettino, M. Pagliai, L. Ciabini, and G. Cardini, J. Phys. Chem. A 105, 11192 (2001). 43. K.A. Yagotintsev, M.A. Strzhemechny, Yu.E. Stetsenko, I.V. Legchenkova, and A.I. Prokhvatilov, Physica B 381, 224 (2006). 44. M. Gu and T.B. Tang, J. Appl. Phys. 93, 2486 (2003). 45. R.C. Yu, N. Tea, M.B. Salamon, D. Lorents, and R. Malhotra, Phys. Rev. Lett. 68, 2050 (1992). 46. F. Negri, G. Orlandi, and F. Zerbetto, Chem. Phys. Lett. 144, 31 (1988). 47. B. Tycko, R.C. Haddon, G. Dabbagh, S.H. Glarum, D.C. Douglass, and A.M. Mujsce. J. Phys. Chem. 95, 518 (1991). 48. M.A. Ivanov and V.M. Loktev, Fiz. Nizk. Temp. 19, 618 (1993) [Low Temp. Phys. 19, 442 (1993)]. 49. A.S. Bakai, Fiz. Nizk. Temp. 32, 1143 (2006) [Low Temp. Phys. 32, 868 (2006)]. 50. M.I. Bagatskii, V.G. Manzhelii, M.A. lvanov, P.I. Muromtsev, and I.Ya. Minchina, Fiz. Nizk. Temp. 18, 1142 (1992) [Sov. J. Low Temp. Phys. 18, 801 (1992)]. 51. V.G. Manzhellii, M.I. Bagatskii, I.Ya. Minchina, and A.N. Aleksandrovskii, J. Low Temp. Phys 111, 257 (1998). 52. F. Gugenberger, R. Heid, C. Meingast, P. Adelmann, M. Braun, H. Wühl, M. Haluška, and H. Kuzmany, Phys. Rev. Lett. 69, 3774(1992). 53. M.A. White, C. Meingast, W.I.F. David, and T. Matsuo, Solid State Commun. 94, 481 (1995). 54. N.I. Odina and A.I. Korobov, Fiz. Tverd. Tela 56, 811 (2014) [Phys. Sol. State 56, 844 (2014)]. Low Temperature Physics/Fizika Nizkikh Temperatur, 2015, v. 41, No. 8 819 Introduction Experiment Results and discussion Conclusion Acknowledgments