Isotopic and spin-nuclear effects in solid hydrogens (Review Article)

The multiple isotopic family of hydrogens (H₂, HD, D₂, HT, DT, T₂) due to large differences in the de Boer quantum parameter and inertia moments displays a diversity of pronounced quantum isotopic solid-state effects. The homonuclear members of this family (H₂, D₂, T₂) due to the permutation symmetr...

Повний опис

Збережено в:
Бібліографічні деталі
Дата:2017
Автори: Freiman, Y.A., Crespo, Y.
Формат: Стаття
Мова:English
Опубліковано: Фізико-технічний інститут низьких температур ім. Б.І. Вєркіна НАН України 2017
Назва видання:Физика низких температур
Онлайн доступ:http://dspace.nbuv.gov.ua/handle/123456789/175360
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:Isotopic and spin-nuclear effects in solid hydrogens (Review Article) / Y.A. Freiman, Y. Crespo // Физика низких температур. — 2017. — Т. 43, № 12. — С. 1687-1706. — Бібліогр.: 97 назв. — англ.

Репозитарії

Digital Library of Periodicals of National Academy of Sciences of Ukraine
id irk-123456789-175360
record_format dspace
spelling irk-123456789-1753602021-02-01T01:29:18Z Isotopic and spin-nuclear effects in solid hydrogens (Review Article) Freiman, Y.A. Crespo, Y. The multiple isotopic family of hydrogens (H₂, HD, D₂, HT, DT, T₂) due to large differences in the de Boer quantum parameter and inertia moments displays a diversity of pronounced quantum isotopic solid-state effects. The homonuclear members of this family (H₂, D₂, T₂) due to the permutation symmetry are subjects of the constraints of quantum mechanics which link the possible rotational states of these molecules to their total nuclear spin giving rise to the existence of two spin-nuclear modifications, ortho- and parahydrogens, possessing substantially different properties. Consequently, hydrogen solids present an unique opportunity for studying both isotope and spin-nuclear effects. The rotational spectra of heteronuclear hydrogens (HD, HT, DT) are free from limitations imposed by the permutation symmetry. As a result, the ground state of these species in solid state is virtually degenerate. The most dramatic consequence of this fact is an effect similar to the Pomeranchuk effect in ³He which in the case of the solid heteronuclear hydrogens manifests itself as the reentrant broken symmetry phase transitions. In this review article we discuss thermodynamic and kinetic effects pertaining to different isotopic and spin-nuclear species, as well as problems that still remain to be solved. 2017 Article Isotopic and spin-nuclear effects in solid hydrogens (Review Article) / Y.A. Freiman, Y. Crespo // Физика низких температур. — 2017. — Т. 43, № 12. — С. 1687-1706. — Бібліогр.: 97 назв. — англ. 0132-6414 PACS: 67.80.ff http://dspace.nbuv.gov.ua/handle/123456789/175360 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
description The multiple isotopic family of hydrogens (H₂, HD, D₂, HT, DT, T₂) due to large differences in the de Boer quantum parameter and inertia moments displays a diversity of pronounced quantum isotopic solid-state effects. The homonuclear members of this family (H₂, D₂, T₂) due to the permutation symmetry are subjects of the constraints of quantum mechanics which link the possible rotational states of these molecules to their total nuclear spin giving rise to the existence of two spin-nuclear modifications, ortho- and parahydrogens, possessing substantially different properties. Consequently, hydrogen solids present an unique opportunity for studying both isotope and spin-nuclear effects. The rotational spectra of heteronuclear hydrogens (HD, HT, DT) are free from limitations imposed by the permutation symmetry. As a result, the ground state of these species in solid state is virtually degenerate. The most dramatic consequence of this fact is an effect similar to the Pomeranchuk effect in ³He which in the case of the solid heteronuclear hydrogens manifests itself as the reentrant broken symmetry phase transitions. In this review article we discuss thermodynamic and kinetic effects pertaining to different isotopic and spin-nuclear species, as well as problems that still remain to be solved.
format Article
author Freiman, Y.A.
Crespo, Y.
spellingShingle Freiman, Y.A.
Crespo, Y.
Isotopic and spin-nuclear effects in solid hydrogens (Review Article)
Физика низких температур
author_facet Freiman, Y.A.
Crespo, Y.
author_sort Freiman, Y.A.
title Isotopic and spin-nuclear effects in solid hydrogens (Review Article)
title_short Isotopic and spin-nuclear effects in solid hydrogens (Review Article)
title_full Isotopic and spin-nuclear effects in solid hydrogens (Review Article)
title_fullStr Isotopic and spin-nuclear effects in solid hydrogens (Review Article)
title_full_unstemmed Isotopic and spin-nuclear effects in solid hydrogens (Review Article)
title_sort isotopic and spin-nuclear effects in solid hydrogens (review article)
publisher Фізико-технічний інститут низьких температур ім. Б.І. Вєркіна НАН України
publishDate 2017
url http://dspace.nbuv.gov.ua/handle/123456789/175360
citation_txt Isotopic and spin-nuclear effects in solid hydrogens (Review Article) / Y.A. Freiman, Y. Crespo // Физика низких температур. — 2017. — Т. 43, № 12. — С. 1687-1706. — Бібліогр.: 97 назв. — англ.
series Физика низких температур
work_keys_str_mv AT freimanya isotopicandspinnucleareffectsinsolidhydrogensreviewarticle
AT crespoy isotopicandspinnucleareffectsinsolidhydrogensreviewarticle
first_indexed 2025-07-15T12:36:51Z
last_indexed 2025-07-15T12:36:51Z
_version_ 1837716484692901888
fulltext Low Temperature Physics/Fizika Nizkikh Temperatur, 2017, v. 43, No. 12, pp. 1687–1706 Isotopic and spin-nuclear effects in solid hydrogens (Review Article) Yuri A. Freiman B. Verkin Institute for Low Temperature Physics and Engineering of the National Academy of Sciences of Ukraine 47 Nauky Ave., Kharkiv 61103, Ukraine yuri.afreiman@gmail.com Yanier Crespo International Center for Theoretical Physics (ICTP), IT-34014 Trieste, Italy E-mail: yanier@hotmail.com Received April 2, 2017, published online October 25, 2017 The multiple isotopic family of hydrogens (H2, HD, D2, HT, DT, T2) due to large differences in the de Boer quantum parameter and inertia moments displays a diversity of pronounced quantum isotopic solid-state effects. The homonuclear members of this family (H2, D2, T2) due to the permutation symmetry are subjects of the con- straints of quantum mechanics which link the possible rotational states of these molecules to their total nuclear spin giving rise to the existence of two spin-nuclear modifications, ortho- and parahydrogens, possessing sub- stantially different properties. Consequently, hydrogen solids present an unique opportunity for studying both isotope and spin-nuclear effects. The rotational spectra of heteronuclear hydrogens (HD, HT, DT) are free from limitations imposed by the permutation symmetry. As a result, the ground state of these species in solid state is virtually degenerate. The most dramatic consequence of this fact is an effect similar to the Pomeranchuk effect in 3He which in the case of the solid heteronuclear hydrogens manifests itself as the reentrant broken symmetry phase transitions. In this review article we discuss thermodynamic and kinetic effects pertaining to different iso- topic and spin-nuclear species, as well as problems that still remain to be solved. PACS: 67.80.ff Molecular hydrogen and isotopes. Keywords: isotopic family of hydrogens, homo- and heteronuclear species, ortho- and parahydrogens, isotope and spin-nuclear effects, ortho- and para conversion, Pomeranchuk effect, reentrant broken symmetry phase transition. Contents 1. Introduction ............................................................................................................................................1688 2. Ortho-para conversion ............................................................................................................................1689 2.1. Experiment .....................................................................................................................................1689 2.2. Theoretical considerations ..............................................................................................................1690 2.3. Conversion in solid D2 ...................................................................................................................1690 2.4. Validity of the concept of ortho-para states at high pressure ..........................................................1691 3. Thermodynamic properties of J = 0 – J = 1 mixtures .............................................................................1692 3.1. Molar volume .................................................................................................................................1692 3.2. HCP lattice distortion in solid H2 as a function of o-p composition ...............................................1692 3.3. Experimental determination of the crystal field splitting of isolated = 1J impurities in solid parahydrogen .................................................................................................................................1694 3.4. HCP lattice distortion in solid HD ..................................................................................................1695 4. Broken symmetry phase transition in solid HD and o-p mixtures ...........................................................1695 4.1. Introduction .....................................................................................................................................1695 4.2. Experimental results ........................................................................................................................1697 © Yuri A. Freiman and Yanier Crespo, 2017 Yuri A. Freiman and Yanier Crespo 4.3. Theoretical studies .......................................................................................................................... 1698 4.3.1. Quantum orientational melting in all-J system ........................................................................ 1698 4.3.2. The reentrant behavior from numerical simulations .................................................................. 1700 4.4. Reentrant phase transitions in ortho-para mixtures ......................................................................... 1701 5. Conclusion .............................................................................................................................................. 1704 References ................................................................................................................................................... 1705 1. Introduction The large isotopic family of hydrogens (H2, HD, D2, HT, DT, T2) presents an unique possibility for studying the diversity of quantum isotopic effects [1–3]. The differ- ences in properties cannot be, as a rule, related solely to the de Boer quantum parameter, since symmetry-related nu- clear spin effects turn out to be far more essential. The requirements of quantum mechanics on a homo- nuclear hydrogen species (H2, D2, T2) rigidly link the rota- tional momentum J and the total nuclear spin molI of the molecule. According to the principle of the indistinguish- ability of identical particles, the molecular wave function of any homonuclear diatomic molecule must be symmetric or antisymmetric under nuclear exchange. The nucleus of hydrogen atom is fermion with nuclear spin = 1/ 2NI and the nucleus of deuterium atom is boson with nuclear spin = 1NI . So the total molecular wave function, which is a product of vibrational, spin, and rotational wave functions, mol vib spin rot=ψ ψ ψ ψ , must be antisymmetric for the hyd- rogen molecule and symmetric for the deuterium molecule. The vibrational part, vibψ , is always symmetric. The spin part, spinψ , for the hydrogen molecule is antisymmetric for the singlet state mol = 0I , and symmetric for the triplet state mol = 1I . Hence the rotational part rotψ is symmetric (rotational quantum number = 0, 2, 4, ...J ) for the singlet state and antisymmetric ( = 1, 3, ...J ) for the triplet state. For the deuterium molecule the spin part is symmetric for mol = 2;0I and antisymmetric for mol = 1I . The rotational part rotψ respectively is symmetric ( = 0J or even) for the symmetric spin state mol = 2;0I state and antisymmetric (Jodd) for the mol = 1I spin state. The states with the parity of the largest possible molI value are called ortho while the states with the other parity are para. Transitions between states of different parity (J or molI ), an ultimately quantum process, called ortho-para conversion, are strictly forbidden in a single molecule and is a low-probability process in a con- densed state. This enables treating the ortho- and para- species as rather stable different substances (orthohydro- gen, o-H2, and parahydrogen, p-H2, and their isotopic counterparts p-D2 and o-D2) with their own properties. The equilibrium ortho-para composition in the mixtures can be calculated by using the rotational energy of a dia- tomic rotator rot rot= ( 1)E B J J + , where rotB is the rota- tional constant (85.25 K for H2 and 42.97 K for D2 [3]), and assuming a Boltzmann distribution. At high tempera- ture the equilibrium ratio of ortho- to para species is de- termined by their statistical weights and is therefore 3:1 for hydrogen and 2:1 for deuterium. Such mixtures are called normal (n-H2 and n-D2). The thermodynamic equilibrium ortho-para concentration for H2 and D2 is shown in Fig. 1. Thus the constraints of quantum mechanics link the possi- ble rotational states of the hydrogen molecule to its total nuclear spins. This symmetry-related link imparts large differences in the solid-state properties of ortho- and para modifications and may be considered as one of the most striking macroscopic manifestations of the microscopic laws of quantum mechanics. Due to this unique relation between the rotational and spin quantum numbers of homonuclear molecules the transitions between J -even and J -odd rotational states with = 1, 3, ...J∆ ± ± require the simultaneous change mol = 1I∆ ± of the total nuclear spin of the molecule. Resulting transformation of the two species into each other is called ortho-para conversion. In isolated molecules the probability of such transitions is negligible and they may be considered as rigorously for- bidden. The conversion in solid hydrogen and deuterium will be considered in Sec. 2. In contrast to H2 and D2, the nuclei in the heteronuclear HD molecule are distinguishable, and no symmetry re- quirements exist on the nuclear wave functions of HD. As a result, the heteronuclear molecules do not have ortho- para species, and for both possible total nuclear spin mol = 1/ 2, 3 / 2I all angular momentum states = 0, 1, 2, ...J and transition between them are allowed. Fig. 1. Equilibrium ortho-para concentration of noninteracting hydrogen and deuterium as a function of temperature from Ref. 1. 1688 Low Temperature Physics/Fizika Nizkikh Temperatur, 2017, v. 43, No. 12 Isotopic and spin-nuclear effects in solid hydrogens 2. Ortho-para conversion 2.1. Experiment While ortho-para conversion is strictly forbidden for isolated molecules, conversion does occur in solid. The ortho-para conversion in solid H2 was first studied experi- mentally as far back as in the beginning of the 1930s by Cremer and Polany [4,5] soon after the discovery of the two hydrogen modifications. At ambient pressure ortho-para conversion in the solid is very slow, taking weeks for a sample to equilibrate [1,2]. It was found that the conversion reaction is autocatalytic, that is, the rate of the reaction is determined by the average number of nearest o-H2 neigh- bors M surrounding each o-H2 molecule: / =dx dt Mx− , where x is the o-H2 concentration. As a result, for the normal equilibrium distribution the rate equation describ- ing the time dependence of the ortho concentration is 2/ =dx dt kx− , (1) where k is the macroscopic conversion constant. The ana- lytic solution of this equation is 1 1 0 =x x kt− −− , which can be used for comparison with experimental results. There have been numerous experimental studies of the ortho-para conversion rate at pressures below 1 GPa by using various methods [6–12]. Ahlers [7] determined the conversion rate by measurement of the heat conduction of a gaseous sample before and after solidification. Pedroni et al. [8] and Schmidt [9] used nuclear magnetic resonance (NMR): the ortho molecules with mol 1= = 1I I contribute to the nuclear magnetization and thus to the NMR signal, while the para molecules with mol 0= = 0I I do not. Silvera’s group [6,11,12] used Raman measurements. The ortho concentration was determined by comparing the rela- tive intensities 0I and 1I of the respective Raman-active rotational transitions = 0 2J → and = 1 3J → . At low temperatures << 6Bk T B , the respective equation takes the form [13] 1 0 1 5 5= / . 3 3 x I I −  +    (2) The experimental data on the conversion rate as a func- tion of density at small compressions up to 0/ 1.7ρ ρ ≈ from Refs. 6–12 are shown in Fig. 2. Here 0ρ is zero- pressure density and ρ is the density at pressure P . The experimental results are seen to be mutually consistent over the whole pressure range. The most extensive measurements at zero pressure were made by Schmidt by NMR at three temperatures 12, 4.2, and 1.57 K in long-term experiment up to 900 h. He gets a conversion rate of = (1.90 0.03) % /k h± which is in agree- ment with most of the previous experiments. Measure- ments at small pressures showed that the conversion rate monotonously increases with pressure. Driessen et al. [6] found a maximum conversion rate at 0.5 GPa followed by a drop up to 0.7 GPa (Fig. 2). Phase diagram studies at high pressures have inspired a renewed interest in the ortho-para conversion in solid H2. Hemley et al. [13] using Raman method raised the pressure limit in the conversion measurements up to 58 GPa. Several characteristic low-temperature (10 K) spectra at 21.6 GPa are given in Fig. 3. They show the decrease in the integrat- ed intensity of the 0 (1)S peak with respect to the 0 (0)S peak with time (here the roton bands 0 (0)S and 0 (1)S in- volve transitions = 0 = 2J J→ , = 1 = 3J J→ , respec- tively). A rather intricate non-monotonous pressure de- pendence of the conversion rate emerged from these studies (Fig. 4). Following maximum at about 0.5 GPa the conversion slows to a minimum at about 3 GPa nearly a factor of two below the ambient pressure rate constant. At higher pres- sures the conversion rate increases rapidly to about 260%/h at 58 GPa. The results are compared with NMR measure- ments to 12.8 GPa [15] ( 0/ 3.7ρ ρ ≈ ). On the whole, there is an excellent agreement between results obtained by the Fig. 2. The conversion rate in solid H2 as a function of the re- duced density at small compressions (after Ref. 6). Fig. 3. Changes in the Raman spectra of solid H2 at 21.6 GPa and 10 K with time elapsed from cool-down completion (after Eggert et al. [14]). Low Temperature Physics/Fizika Nizkikh Temperatur, 2017, v. 43, No. 12 1689 Yuri A. Freiman and Yanier Crespo two methods with possible discrepancies arising from tem- perature differences in the two studies. 2.2. Theoretical considerations A theory of the ortho-para conversion in solid hydrogen was first developed by Motizuki and Nagamiya [17]. They showed that the conversion rate is controlled by the three mechanisms: (1) the magnetic dipole-dipole interaction of the nuclear spins of ortho molcules; (2) magnetic-dipole in- teraction between the nuclear spins and the rotational mag- netic moments; (3) magnetic interaction with paramagnetic impurities. The conversion transition ratio W can be calculated from the golden rule: 2 int 2= | | | | ( ),i i f if W P f i E Eπ 〈 〉 δ −∑   (3) where iP is the probability that the system is initially in one of the states i , and f is the final state; iE and fE are the initial and final energies of the system. int is the in- teraction Hamiltoman and has three additive parts which can change the nuclear spin states: ss , the nuclear spin- spin interaction; rs , the rotation-spin interaction; and qq , the electric quadrupole interaction, (explicit equations can be found in Refs. 3, 18). ss is the dominant term with rs yielding a 2–3% correction to ss . The delta-function in Eq. (3) requires conservation of energy. The rotational ener- gy change in a conversion is rot ( 1)B J J + . According to the theoretical description of ortho-para conversion in solid hydrogen as developed by Motizuki and Nagamiya [17], an ortho-para transition requires a large magnetic field gradi- ent acting on a molecule to flip a proton spin. This field gradient is supplied by the rotational magnetic moment of neighboring ortho molecules. Thus two ortho-neighbors are needed for one of them to convert. As a result, if the ortho distribution is random, the rate equation Eq. (1) de- scribes the time dependence of the ortho concentration. The rotational energy difference between a = 1J ortho molecule and a = 0J para molecule, 0 1 rot= 2 =E B←∆ = 171 K, released by ortho-para conversion must be ab- sorbed by the lattice. The Debye temperature DΘ gives an estimate for the limiting phonon energy. Since at ambient pressure DΘ = 118.5 K [3] the energy conservation re- quires the simultaneous creation at least two phonons. The two-phonon theory of Motizuki and Nagamiya predicts a drop in the conversion rate with increasing pressure over about 1 kbar where the two-phonon conversion rate is neg- ligible. Thus, neither the two-phonon theory nor the one- phonon theory in the harmonic approximation can explain experimental results [7,8,10] which showed the conversion rate to be an increasing function of density. Therefore Ber- linsky [19] suggested that the effect of the large anhar- monisity in solid hydrogen is to broaden the high-energy features of the phonon spectrum into a high-energy tail which extends far above B Dk Θ . On the basis of this con- jecture Berlinsky [19] developed a detailed theory of one- phonon processes and found that the one-phonon con- version rate should initially increase with pressure before falling above about 0.5 GPa because of density of states factors and the rapid increase in phonon energy. On the base of the conversion rate data [14,15] a new conversion mechanism was proposed by Strzhemechny and Hemley in Ref. 16 and by Strzhemechny et al. in Ref. 20 that differs from that employed to explain low-pressure data. The steep increase in conversion rates at high pres- sures was explained by a conversion channel that involves an intermediate state in which new excitations are created due to the electric quadrupole-quadrupole (EQQ) coupling between rotational momenta. The enhancement was explain- ed by a gap closing that arises when the EQQ interaction becomes sufficient to substantially diminish the conversion energy released because of the lowering of the ground- state level of the converted ortho molecule. It was shown that this concentration-sensitive channel comes into play at compressions 0= /ξ ρ ρ between 3 and 4 and ceases to ope- rate at higher values ( 6–7ξ ≈ ), depending on the ortho concentration. Thus, this theory predicts a significant re- duction in the conversion rate at about 80 GPa; The exten- sion of the measurements to pressures over 50 GPa would provide a critical test of the conversion mechanisms pro- posed in Refs. 16, 20. 2.3. Conversion in solid D2 Theory of conversion in solid D2 was developed first by Motizuki [21]. The conversion in D2 differs from that in H2 in a number of essential aspects. First of all, the deuter- ium nucleus, deuteron, possesses a nonzero quadrupole Fig. 4. The conversion rate vs. reduced density (the crystal densi- ty ratio reduced to the = 0P value) at high compression. Solid circles are from Refs. 13, 14, empty squares are from Ref. 15. The solid curve is the theoretical prediction for hcp n-H2 from Ref. 16. The open circles are earlier data (see Fig. 2). Inset: Low density region (see Ref. 14). After Eggert et al. [14]. 1690 Low Temperature Physics/Fizika Nizkikh Temperatur, 2017, v. 43, No. 12 Isotopic and spin-nuclear effects in solid hydrogens moment. The interaction of this quadrupole moment with electric field gradients due to neighboring molecules pro- vides another mechanism for conversion in addition to dipolar interactions. Also, the deuteron has = 1DI , so the final state of the = 0J D2 molecule may be any of the six nuclear states corresponding to = 0I or 2. If the final nu- clear state is = 2I , then this molecule will cause magnetic field gradients at neighboring molecules, and thus it may still contribute to the conversion process. This is in con- trast to the case of solid H2 where the para molecules have = 0J , mol = 0I and thus do not contribute to further con- version. One more significant difference between the conversion in D2 and H2 lies in the fact that the rotational constant of D2, D2 rot = 42.97 KB [3], about a half of that for H2 (85.25 K) with the respective change of the rotational energy in going from the = 1J to the = 0J state, while the Debye tem- perature ( D2 DΘ ) = 114.1 K [3] is nearly the same ( H2 DΘ ) = = 118.5 K [3]. Thus the one-phonon process is allowed for conversion in solid D2. The presence of the conversion mechanism that in- volves = 0J molecules with = 2I changes the rate equa- tion describing the time dependence of the concentration of = 1J molecules: 2/ = (1 ).dx dt kx k x x′− − − (4) The general solution of this equation with the initial condition 0( = 0) =x t x has the form 1 1 1 0 0= ( 1 / )[1 exp ( )].x x x k k k t− − − ′ ′− − + − (5) A characteristic distinction of the conversion in deuteri- um is that at low = 1J fraction the conversion rate is linear in the concentration x of the = 1J molecules. Experi- mental results obtained by different methods (see Table 1) gave rather close results: Table 1. The conversion constants in solid deuterium at = 0P Author, Ref., Method 4 1(10 h )k − − 4 1(10 h )k − −′ Hardy and Berlinsky [22] (NMR) 5.6 =k k′ Milenko and Sibileva [23] 5.6 0.5± 5.3 0.3± Berkhout, Minneboo, and Silvera [24] (Raman) 6.3 0.1± =k k′ Calkins, Banke, Li, and Meyer [25] (NMR) 5.5 0.5± =k k′ Bagatskii, Krivchikov, Manzhelii et al. [26] 6.4 0.2± Strzhemechny and Tokar [27] (Theory) 6.1 14.2 Comparison with the conversion rate in H2 ( 2 11.9 10 h− −⋅ ) shows that in D2 it proceeds about 30 times slower. The efficiency of various conversion channels in solid deuterium at high pressures have been considered by Strzhemechny and Hemley [28]. They found that the standard phonon- assisted channels are inefficient at high pressures in D2 as they are in H2, and the idea of the intermediate state with subsequent participation of the EQQ interaction remains productive for D2. 2.4. Validity of the concept of ortho-para states at high pressure At small pressures when the wave function of the crys- tal may be represented as a product of molecular wave functions there are no any doubts that the notion of ortho- para states is a well justified quantum-mechanical concept. To what extent this concept remains valid at high pressures when the rotational quantum number J is no longer a good quantum number? How does an essential increase in the conversion rate with pressure affect the ortho-para con- cept? These important questions were analyzed in detail by Silvera [29] and Silvera and Pravica [30]. Their main con- clusion is that the only important quantum number for the o-p states is parity under exchange and that the nuclear spin states preserve up to pressures when the dissociative Wigner-Huntington transition occurs. In solid molecular hydrogen or deuterium the many- body wave function must be symmetrized with respect to exchange of nucleons. Nucleon exchange can take place within a molecule or between molecules. If exchange of nucleons between neighboring molecules took place, then the structure of the wave functions under nucleon permuta- tion would be very complicated and the ortho-para concept associated with isolated molecules would not be applicable. The total molecular wave function can be written as a superposition of pure ortho and pure para states: ortho para=Ψ Ψ +αΨ . (6) For the validity of the concept of ortho-para species, the mixing parameter α should be very small. The perturbation which causes mixing is the same that causes conversion. Taking into consideration the mecha- nism of Motizuki and Nagamiya Silvera obtained the fol- lowing density dependence of the resultant mixing parame- ter [29]: 10 4/3 1 0| | = 8 10 ( / ) .c−α ⋅ ρ ρ (7) where 1c is the ortho concentration. From Eq. (7) we see that at zero pressure the mixing parameter is exceptionally small. The highest compression 0= /ξ ρ ρ reached in the equation of state experiments at 180 GPa is 10.4 [31]. The mixing parameter increases by a factor 22.7 but remains very small of order 810− , and the concept of ortho-para species remains valid to very high pressure. There is an interesting question concerning a possible implication of ortho-para species for metallization [30]. At a finite temperature we have a mixture of ortho- and parastates which violates the translational invariance. This can possibly give rise to Anderson localization and sup- Low Temperature Physics/Fizika Nizkikh Temperatur, 2017, v. 43, No. 12 1691 Yuri A. Freiman and Yanier Crespo press an insulator–metal transition that would occur in a pure lattice [32]. It might be the case that even pure ortho- or para lattices suffer from Anderson localization. Since the electron motions are very much faster than the nuclear motions, an electron moving through a pure lattice sees a snapshot of molecules with random orientations, although on the average (in the quantum mechanical sense), the molecules have well defined distributions. At this time it is not clear from theoretical considerations whether the elec- tron feels a translationally invariant or a disordered lattice. It is possible that the failure to observe metallization in the phase III of hydrogen, above 150 GPa, is due to this type of disorder. 3. Thermodynamic properties of J = 0 – J = 1 mixtures 3.1. Molar volume In 1972 Silvera et al. [33] studying Raman-active 2gE phonon mode in hcp hydrogen and deuterium found unex- pectedly large downward shift ( 5% ) in the measured frequency when going from pure = 1J to pure = 0J solid. This effect was explained by Silvera et al. [12] by the dif- ference in molar volume of = 1J and = 0J species, the former having a smaller molar volume. The free energy has translational (lattice) contributions and orientational contributions related with isotropic and anisotropic interac- tions, respectively. When compressions are not too high, we may restrict ourselves to the contribution from EQQ (electric quadrupole-quadrupole) interactions. The quadru- pole pressure difference between = 1J and = 0J H2 and D2 in the orientationally ordered Pa 3 state is shown in Fig. 5 [11]. Hydrogen is a quantum solid and is therefore highly compressible. Due to the large compressibility, the weak anisotropic forces related with = 1J species have a non-negligible effect on the molar volume. The zero- pressure molar volume V of hydrogen and deuterium as a function of = 1J concentration is shown in Fig. 6. In the orientationally disordered state of hexagonal close packed (hcp) solids at = 4.2 KT we have the following dependen- cies [1]: 2 3 2( ) = 23.16 0.091 0.217 cm / mol (H )V x x x− − , and 2 3 2( ) = 19.95 0.16 0.04 cm / mol (D )V x x x− − . 3.2. HCP lattice distortion in solid H2 as a function of o-p composition At zero pressure the molecules in J -even solid hydrogens (p-H2, o-D2) are virtually spherical. Rigid spheres in the undistorted lattice crystallize into fcc (face center cubic) or hcp lattice. As compared with fcc, the hcp lattice has an additional degree of freedom associated with the /c a ratio. A lattice of closed packed hard spheres has / = 8 / 3 1.633c a ≈ (an ideal hcp structure). The quantity = / 8 / 3c aδ − , the lattice distortion parameter, describes the deviation of the axial ratio from the ideal value. In the case of < 0δ , this distortion involves extension within close-packed planes, and contraction along the c-axis di- rection, and vice versa, for >0δ the lattice is expanded along the c axis and contracted within close-packed planes. Calculations with pair isotropic potentials have shown that the ideal hcp lattice does not correspond to minimal lattice energy [35–37]. Calculations with a many-body potential and DFT calculations performed for solid He showed [38] that the pressure dependencies of the lattice distortion parameter ( )Pδ for a many-body (two- plus three-body) and for pair intermolecular potentials are qualitatively different. The three-body forces flatten the lattice ( < 0δ ) while the pair forces at large compressions tend to elongate it ( > 0δ ). Thus, it was shown that the lat- tice distortion parameter is a thermodynamic characteristic which is very sensitive to the many-body component of the intermolecular potential and can therefore be used as a probe of the many-body forces [39]. The deviation of the axial ratio from the ideal value in the hcp solids can be attributed ultimately to a lowering of the band-structure energy through lattice distortion. In the molecular hcp solids we can explicitly ascribe this effect to the reduction in the ground-state energy due to lattice dis- tortion. In the case of solid hydrogens the effect of lattice distortion both on the isotropic and rotational components of the zero-point energy is essential. The orientational state of molecules in the distorted lat- tice is characterized by the orientational order parameter 20= 4 / 5 Yη π 〈 〉, where (...)〈 〉 means thermodynamic av- eraging with the rotational Hamiltonian. The lattice- rotation coupling is described by the the Hamiltonian int 2 20= 4 / 5c Y−ε π , where 2cε is the crystal-field pa- rameter [2] which is linear with respect to δ : 2 1= ; = 6 , 2c dBB B B R dR  ε δ − +      (8) Fig. 5. The quadrupole pressure in p-H2 and o-D2 in 3Pa ordered state at = 0T (after Jochemsen et al. [11]). 1692 Low Temperature Physics/Fizika Nizkikh Temperatur, 2017, v. 43, No. 12 Isotopic and spin-nuclear effects in solid hydrogens where ( )B R is the radial function of the single-molecular term in the anisotropic intermolecular potential [2]. Thus, the state of the lattice can be described by two coupled order parameters, ( , )P Tη and ( , )P Tδ , which can be found by the minimization of free energy with respect to these parameters. Figure 6 shows the volume dependence of the lattice distortion parameter δ and orientational order pa- rameter η in parahydrogen and orthodeuterium calculated using the many-body intermolecular potential from Ref. 40. The negative sign of δ means that the lattice is flattened compared with the ideal one. For all hcp elemental solids except helium, hydrogen, and high-pressure Ar, Kr, and Xe the behavior of δ with pressure and temperature is well established from both theory and experiment. Typical values are of the order of 10 2− . For solid helium δ is an order of magnitude small- er [38,41]. The first measurements of /c a ratio in solid hydrogens were done by Keesom et al. [42] who found that at zero pressure hcp lattice of p-H2 is close to ideal. X-ray zero-pressure study by Krupskii et al. [43] confirmed this result ( / = 1.633 0.001c a ± ) and extended it for o-D2. In fact, the only structural study of p-H2 and o-D2 at elevated pressures up to 2.5 GPa and low temperatures were made by Ishmaev et al. using the neutron diffraction method [44,45]. It was found that the ratio /c a is practically constant and is slightly less than the ideal hcp value (1.631 0.002± ). In the absence of direct experimental data some qualita- tive conclusions on the lattice distortion parameter of p-H2 were obtained by Goncharov et al. [46] from low-fre- quency Raman spectra at low temperature. The authors have measured low-frequency Raman spectra at low tem- perature for the pressure range up to the I–II phase transi- tion and used these spectra to estimate the crystal-field parameter 2cε . Assuming that only the second-order crystal field is responsible for the splitting of the roton triplet band 0 (0)S and get that 2| |cε  1 cm–1 and thus 3 4| | 10 10− −δ − in accord with theoretical data shown in Fig. 7. There were numerous structural studies of n-H2 and n-D2 [3,31,48–51]. Synchrotron single-crystal x-ray diffraction measurements of n-H2 [31,48] and n-D2 [48] up to mega- bar pressures at room temperature revealed an approxi- mately linear decrease of the /c a ratio with increasing pressure (Fig. 8). No isotope effect in the pressure depend- ence of the /c a ratio was found. Fig. 6. The molar volume at zero pressure as a function of con- centration of = 1J modification (after Driessen et al. [34]). Fig. 7. (Color online) Lattice distortion parameter δ and orien- tational order parameter η in parahydrogen and orthodeuterium as a function of molar volume [47]. Fig. 8. (Color online) Lattice distortion parameter in normal or- tho-para mixture of solid hydrogens. Theory: solid line [47]; ex- periment: red squares — data from Ref. 48, blue triangles — data from Ref. 31. Low Temperature Physics/Fizika Nizkikh Temperatur, 2017, v. 43, No. 12 1693 Yuri A. Freiman and Yanier Crespo For a single = 1J molecule in a lattice of = 0J mole- cules there is an additional contribution to the ground-state energy arising from the polarization of the surrounding = 0J molecules by the the electric quadrupole field of the = 1J molecule. The polarization energy due to the interac- tion of the quadrupole moment of the = 1J molecule with the induced dipole moments of the surrounding nearest neighboring = 0J molecules equal to 2 8 1 = 18 ,Q R−ε − α (9) where α is the polarizability of the = 0J molecules and Q is the quadrupole moment of the = 1J molecule, and R is the intermolecular distance. If the crystal shows a homogeneous deviation from the ideal hcp structure, the polarization energy contains a crys- tal-field term [2] 2 20 4= ( ), 5c cV Yπε Ω (10) where Ω specifies the orientation of the = 1J molecule with 2 1 24= . 7c cε − ε δ (11) The splitting of the = 1J level in the crystal field cV is given by 2 3= ( ) (0) = | |, 5c cE E∆ ± − ε (12) where ( )E M is the energy of the state = 1J , =zJ M with z along the c axis. The positive sign of c∆ implies that the singlet state = 0zJ is the ground state of the triplet. Due to this splitting the ground-state energy is brought down by 22 | | /5cε . Thus, the gain in the anisotropic interaction contribution to the zero-point energy is linear in the lattice distortion parameter δ . At the same time, this distortion increases the contribution to the ground-state energy from the isotropic part of the intermolecular interaction. This contribution is quadratic in the lattice distortion parameter δ [52]. The loss in the isotropic part and the gain in the anisotropic part of the ground-state energy determines the lattice distortion parameter at the given molar volume. One cannot obtain a reliable value of δ from Eqs. (9)–(12) be- cause there are other contributions to the splitting Eq. (12) [2]. Assuming that the volume dependence of the polariza- tion energy holds the same volume dependence as in Eq. (9) in spite of these contributions, we will get the pressure dependence of the lattice distortion parameter shown in Fig. 8. When this result is compared with that for p-H2 (Fig. 7) it is seen that the presence of = 1J molecules in- creases the hcp lattice distortion by two order of magni- tude. 3.3. Experimental determination of the crystal field splitting of isolated = 1J impurities in solid parahydrogen An isolated = 1J orthohydrogen substitutional mole- cule in the pure = 0J hcp crystal lattice is an ideal system which has been studied extensively both theoretically and experimentally [2]. The heat capacity anomaly in the = 0J solid hydrogen samples containing small concentrations of = 1J impurity molecules has been found by Mendesson, Ruhemann and Simon as early as in 1931 [53] and ex- plained by Schaefer [54] in 1939 by the splitting c∆ of the triplet level of the = 1J impurity molecules (Eq. (12)). In succeeding years this splitting has been studying in numer- ous heat capacity [57,58] and NMR [22,63–65] measure- ments, and theoretical works [55,56,59–61,66,67]. The first direct measurements of the crystal field split- ting of the = 1J sublevels were performed by Dickson, Bayers and Oka [68] who observed the fine structure of the infrared 3(0)Q second overtone transition ( = 3 0,ν ← = 0 0)J ← at 11758 cm–1 in parahydrogen crystals con- taining approximately 0.1% orthohydrogen. The transition is caused by the dipole moments induced in = 0J H2 due to the quadrupolar electric field of the = 1J H2. Three main sets of features were observed: the stronger sets separated by nearly 0.5 cm–1 which arise from nearest neighbor pairs of = 1J H2 and = 0J H2, while the weaker set near the center is due to next nearest neighbor pairs. The large splitting of 0.5 cm–1 corresponds to that of the = 1J H2 next to the vibrationally excited = 0J H2: the vibrational excitation breaks the 3hD symmetry of the crys- tal and the splitting is greatly enhanced since the lattice sum for nearest neighbors no longer vanishes. The break- down of symmetry also removes the degeneracy of the = 1M ± levels; this together with the slight difference in energy for in plane and out of plane pairs leads to the intri- cate structure of each set. The small = 1J H2 splitting without vibrational excita- tion for = 0J H2 is shown in Fig. 9 where the set of transi- tions near 11758.39 cm–1 is given with an expanded fre- quency scale for the polarization of the laser radiation both parallel and perpendicular to the crystalline c axis. Alto- Fig. 9. The low-frequency set of transitions for both parallel (up- per) and perpendicular (lower) polarizations. The first and third lowest frequency transitions are due to out-of-plane nearest neigh- bor = 1 / = 0J J hydrogen molecules while the second and fourth are due to in-plane neighbors [68]. 1694 Low Temperature Physics/Fizika Nizkikh Temperatur, 2017, v. 43, No. 12 Isotopic and spin-nuclear effects in solid hydrogens gether eight such pairs of transitions are observed and led to the value = (0.0102 0.0003) Kc∆ + and to the observa- tion that the = 1M ± levels are higher in energy than is the = 0M level. This value agrees with the results of Schweizer et al. [64] within their quoted uncertainty and is close to the theoretical estimate of Raich and Kanney [66]. 3.4. HCP lattice distortion in solid HD The principal difference between the HD molecule and its homonuclear isotopic analog is that the center of mass is considerably shifted relative to the center of electron density distribution. A free HD molecule rotates around its center-of-mass, which is separated from the midpoint be- tween the two nuclei, the center-of-charge, by a distance 1 6 er , where er is the internuclear distance ( = 0.74116 Åer ). Due to such mismatch the heteronuclear solid isotopes HD, HT, and DT have some properties that distinguish them from the homonuclear species. The most remarkable case — the reentrant BSP phase transition in solid HD will be dis- cussed in Sec. 4. In this section we discuss the effect of the molecular asymmetry on the axial ratio of the hcp lattice parameters /c a . Structure data obtained using x-ray diffraction [69] (Fig. 10) gave an unexpected result: c/a = 1.618 ± 0.003 at the zero-temperature limit while for p-H2 x-ray diffraction measurements by Krupskii et al. [43] gave the ideal value 1.633 ± 0.001. Surprisingly, the /c a ratio showed different temperature behavior for different species: whereas in both normal and para H2 the c/a ratio slightly decreased [69] upon warming, the c/a ratio in HD grew with temperature, reaching the value 1.623 ± 0.005 at 17 K. The authors qua- litatively explained the found effect by the different virtual contributions of the excited rotational states into the mole- cular ground rotational state of the molecule. Theory of the effect developed recently by Strzhemechny and Hemley [74] explains how differences in the rotational dynamics affect the value of the /c a ratio at low temperatures. Their approach is based on the following equation: el rot[ ( ) ( )] = 0,E E∂ ∆ δ + ∆ δ ∂δ (13) where elE and rotE are the densities of the elastic and ground-state rotational energies, respectively, as functions of the deviation δ . While the loss in the elastic energy el ( )E∆ δ is proportional to 2δ , the gain in the rotational energy rot ( )E∆ δ due to the quantum self-polarization effect calculated in the second perturbation order in the aniso- tropic interactions is linear in δ . Taking both contributions into consideration one gets a nonzero deviation δ . The final estimate for δ in solid HD is 3(HD) = (0.61 0.08) 10 ,−δ − ± ⋅ (14) the deviation δ in HD is also negative and approximately 50 times larger in magnitude than in H2 or D2 at zero pres- sure (see Fig. 7). Though the above value is by a factor of approximately 20 smaller compared with what was deter- mined experimentally [69], the trend is correct. 4. Broken symmetry phase transition in solid HD and o-p mixtures 4.1. Introduction To a very good approximation, the electron density dis- tribution in the H2 and HD molecules are the same, but in the HD molecule the center-of-charge does not coincide with the center-of-mass. Since the molecule rotates around its center-of-mass but the intermolecular interaction is re- lated to the center-of-charge molecule, rotations of the mo- lecules are accompanied by translational displacements of the center-of-mass. Thus, rotation and translation of the molecule are coupled dynamically. As a result of such off- center rotation, an additional Heisenberg-like terms appear in the anisotropic part of the intermolecular potential [2,71]. Evidence for differences in properties of symmetric and asymmetric hydrogens resulted from the rotational-trans- lational coupling in the condensed state has been reported since the sixties of the last century [72]. The heteronuclear hydrogen molecules have nonzero static electric dipole moments. The experimental value of the HD dipole moment, as derived from pure rotational spectra of gaseous HD, is 4= 5.85 10p −⋅ D [73]. Because of the nonequivalency of the ends of the HD molecule, there is an additional dynamics in solid HD in comparison with homonuclear isotopes H2 and D2, namely, end-to-end reorientational dynamics [2]. In principle, by ordering the dipole electric moments it is possible to minimize the di- pole-dipole interaction energy. Thermodynamics and ki- netics of the dipole ordering have been discussed in litera- ture in connection with the problem of the dipolar ordering and zero-temperature entropy in solid CO, a diatomic pos- sessing in addition to a large quadrupole moment a small permanent dipole moment 0.1 D, two orders of magni- tude larger than HD [75]. In the case of solid CO the tem- Fig. 10. Temperature dependencies of the axial ratio of the hcp lattice parameters c/a for deuterohydrogen [69], parahydrogen [43], and normal hydrogen [70]. After Ref. 69. Low Temperature Physics/Fizika Nizkikh Temperatur, 2017, v. 43, No. 12 1695 Yuri A. Freiman and Yanier Crespo perature of the hypotetical dipole ordering phase transition is about 5 K [3], but the transition is frozen because the molecules are ordered by much stronger quadrupole- quadrupole interaction at T  40 K. The qualitative dis- cussion of the possible dipole ordering in solid HD was done by Chijoke and Silvera [76]. The characteristic dipo- lar energy of an HD molecule in the field of the nearest neighbors 2 3/dE zp R (where z is the number of nearest neighbors and nnR is the nearest neighbor distance) gives for the transition temperature an estimate 1 mK. Such transition cannot take place in the quantum phase I since all the molecules are in the = 0J state and the expectation value of the dipole moment in this state is zero. It has been speculated by the authors of Ref. 76 that the conditions to observe the dipole ordering transition in HD are much more favorable at high pressure. They estimated that at P  150 GPa the dipole moment induced in the molecule by the quadrupole moment of neighboring molecule in = 1J state is 24 times larger than the permanent dipole moment and this could increase the interaction energy to tens of K. Summarizing results of the IR study Ref. 76 the authors stated that in their low-temperature spectra they “found no indication of a transition of electric dipole order- ing”. We think that even though their considerations about the strong enhancement of the dipolar interaction at high pressure are correct, the dipolar ordering in HD is inobserv- able due to the same reason why it cannot be realized in solid CO: dipole end-to-end reorientations are frozen in a strong quadrupolar field. The BSP transitions in the even-J species (parahydrogen and orthodeuterium) and all-J system (HD) provide an ex- ample of quantum phase transitions where the orientational ordering (OO) is realized by a competition between the potential and kinetic energies of the system as opposed to the classical phase transitions in the J -odd species (ortho- hydrogen and paradeuterium) where the OO is realized by a competition between the potential energy and the en- tropy contribution to the free energy of the system. As a re- sult, strong differences are observed in the OO properties of odd-J, even-J, and all-J (e.g., HD) systems. The odd-J systems order at any pressure at low temperature, while the molecules in the even- and all-J systems at low pressure are in the = 0J quantum state and thus are spherical enti- ties. The OO in the even- and all-J systems is realized by increasing the pressure, that is by increasing the potential energy of the molecules. By analogy with quantum melting of solid helium the quantum orientational disordering in even- and all-J solid hydrogens was taken the title Quan- tum Orientational Melting (QOM) [80–83]. The phase diagram of all-J systems possesses a surprising anomalous feature: there is a pressure range 0< <mP P P 0(P is the ordering pressure at = 0T , mP is a minimal or- dering pressure at the transition line) where the disordered phase is reentrant (see Fig. 11). At = 0T in this pressure range the system is in the disordered state but on heating we cross the phase transition line and get to the ordered phase. By increasing the temperature further at constant P we get once again to the disordered phase. And conversely, if we start from the ordered state inside the pressure range 0< <mP P P , we can get from the ordered to the disordered state by two ways: either by heating or by cooling! In the former case the order is destroyed by thermal fluctuations, in the latter case the order is destroyed by quantum fluctua- tions. This reentrant behavior was first predicted for the all-J system by mean-field calculations [80,81] and found experimentally in HD by Silvera’s group [78]. As was shown in Refs. 80, 81, there is a very close analogy between the behavior of the melting line in solid 3He and the BSP transition in HD. As known, at the low- temperature section of the melting line of 3He, the entropy of liquid is smaller than that of the solid (Pomeranchuk effect). As a result, there is a minimum in the pressure- temperature dependence of the melting line in 3He — the melting transition in 3He is reentrant. It was shown that the reentrance in the BSP transition line in all-J rotor systems has also entropy-driven origin [80,81]. Path integral Monte Carlo (PIMC) technique for studying the reentrant behavior in solid HD was applied in Ref. 84. The system of asymmetric rotors was considered with cen- ters fixed on fcc lattice with the EQQ interaction as the anisotropic interaction potential. The results were only of qualitative value. Recently a more accurate approach was implemented by Crespo et al. [85] where the authors ap- plied to solid HD at high densities PIMC within the con- stant pressure ensemble. They considered both the transla- tional and rotational degrees of freedom and different sources for the anisotropic interaction potential. The HD phase line was calculated for both fcc and hcp lattices, dis- playing the reentrant behavior in both cases. Good agree- ment with the experimental data was obtained in the case of the hcp lattice, including the minimum pressure ( mP ) where the BSP transition occurs ( mP = 56 GPa) and tem- perature at the minimum point ( mT = 25 K) to be compared with experiment: mP = 53 GPa, mT = 30 K [78]. Fig. 11. Schematic experimental phase diagram for H2, ortho-D2 and HD from data reported in Refs. 77–79. 1696 Low Temperature Physics/Fizika Nizkikh Temperatur, 2017, v. 43, No. 12 Isotopic and spin-nuclear effects in solid hydrogens 4.2. Experimental results Experimentally, “Remarkable high pressure phase line of orientational order in solid hydrogen deuteride” (as was the title of the first publication), was first found by means of Raman scattering by Silvera’s group [78] and then studied to higher pressure by using near infrared spectroscopy [76]. It was observed that in HD the BSP transition pressure first decreases and then increases with increasing temperature, a very unusual reentrant behavior, different from the homo- nuclear isotopes H2 and D2. In this Raman study the authors measured both the Raman active rotons and vibrons in HD to pressures of 120 GPa and temperatures as low as 3.5 K (Fig. 12(a) [78]). The HD vibron frequency shifts substan- tially with changing pressure at the phase line and due to the sharpness of the vibron the detection of this discontinu- ity was used as a method for determining the phase line. In Fig. 12(b) the discontinuity is seen on the 64 GPa curve. The intensity of the peak shifts from one vibron to the other when the phase line is crossed. At the lowest temperature of this experiment (3 K) a critical pressure = 68. Pa3 GcP GPa was obtained. The critical pressure at = 0T was esti- mated by extrapolation to be 0 = (69.0 2) GPaP ± , lying between the values for H2 and D2 solids. As the Raman line broadened with increasing pressure and the shift of the line became too small to detect the BSP, the work Ref. 78 was limited in pressure by 74 GPa. Another technique that can be used to study this phase transition at low temperatures is the infrared (IR) absorp- tion. In Ref. 76 the authors reported the first infrared ab- sorption measurements on HD to pressures of 159 GPa. In Fig. 13 the IR spectra are shown in the region of the first- order band for pressures between 11 and 159 GPa at 5 K. These spectra correspond to the (0 1)v → vibron transition. The 0 1→ vibrons of HD in hcp solid are symmetry for- bidden for IR absorption at low temperature when the thermal occupation of the = 1J level is close to zero. The transition to the BSP phase was determined by the appear- ance of two strong and relatively sharp absorption peaks with linewidths of 18 and 40 cm–1 (see the curve at 77 GPa in Fig. 13), as well as very weak third peak, indicated by arrows in Fig. 13. These lines were interpreted as IR- allowed v (0 1→ ) vibrons in the BSP phase. The reentrant behavior of the BSP phase was observed by the appearance and disappearance of these lines as the temperature was raised at a fixed pressure of 56 GPa. In the ultra-high- pressure region (159 GPa) it was found that the system undergoes a transition from the BSP to the phase III. At pressures of (157 ± 3) GPa it was observed the appearance of a new spectral feature in the IR spectrum, a new peak at 3890 cm–1 (see the arrow in 159 GPa in Fig. 13). This is Fig. 13. Absorption spectra for several pressures, collected at temperatures of 4–8 K (after Ref. 78). Fig. 12. Raman spectra of HD for several pressures with tempera- ture 5–7 K. (a) Rotational spectra, (b) vibron spectra (after Ref. 78). Low Temperature Physics/Fizika Nizkikh Temperatur, 2017, v. 43, No. 12 1697 Yuri A. Freiman and Yanier Crespo the same pressure range and spectral response for the onset of the phase III in H2 [86] and D2 [87]. As a result, these data were were interpreted as a first observation of phase III in HD [76]. In Fig. 14 the absorption spectra are shown collected at 159 GPa for several temperatures. As can be seen from Fig. 14, the intensity of the peaks corresponding to both the phase III and BSP decreases as temperature is raised becoming at 120 K the characteristic spectrum of the hcp disordered phase. The phase transition line determined from these methods are shown in Fig. 15. 4.3. Theoretical studies 4.3.1. Quantum orientational melting in all-J system The BSP transition in solid hydrogen, is a realistic ex- ample of Quantum Orientational Melting (QOM) that can occur at zero temperature. This quantum phase transition is driven by a competition between the potential energy, which tends to order the system, and the kinetic energy, which tends to delocalize the system. Early mean field theore- tical models have shown that this type of transition can be observed in a system of rotors under appropriate condi- tions [80–82]. The mean field approximation considers a system of N linear quantum rotors described by the Hamil- tonian [80–82] 2 2 rot 20 ˆ= ( ) / 2, N N i i i B L U Y UN− η Ω + η∑ ∑ (15) where L̂ is the angular momentum operator, U is a mo- lecular field constant (U P ), rotB is the rotational con- stant = 4 / 5,η π 20 ( )iY〈 Ω 〉 is the order parameter, and < > denotes thermodynamic averaging with the Hamil- tonian (15). The computational scheme used to find the spectrum of the Hamiltonian (15) is given in Refs. 80–82, 90. Figure 16 shows the energy spectrum for a linear rotor in the field 2 (cos ( ))VP− ϑ where =V Uη. As can be seen from this fi- gure, in the disordered phase ( = 0η ) the gap ∆ between the ground state and the first exited triplet state is rot= 2B∆ . In the ordered phase of an all-J system, instead, this degeneracy is splitted into a singlet and a doublet so that ∆ decreases as increasing η (for constant U ). In the the strong interaction limit the ground state becomes a doublet. This reduction of the gap allows the occupation of this level even at low temperatures and as a result there is an additional contribution to the entropy, of the order of ln 2Bk per a molecule where Bk is the Boltzmann con- stant, that stabilizes the ordered phase at low temperatures. The previous reasoning was confirmed by the calcula- tions of the entropy and order parameter with temperature. Figure 17 shows the temperature dependence of the entro- py of both the one corresponding to the free rotor and that of the OO state for different values of /U B (11.5 < / < 15.0U B ). As can be seen, in the low-T region the entropy of the OO phase is larger than that of the dis- ordered one. The inset shows the temperature dependence of the difference in entropy ( 0=S S S∆ − ) between the or- dered, S , and the disordered, 0S , phases for / 12.9U B ≈ . The free energy of the system as a function of the order parameter ( )η was calculated using the obtained spec- Fig. 14. Absorption spectra collected at 159 GPa for several tem- peratures in the range between 5–160 K (after Ref. 76). Fig. 15. Phase diagram of HD, the phase lines determined in Ref. 76 are shown as solid lines and the phase line determined earlier by Raman measurements as dashed lines [78]. The phase lines of D2 (dotted lines) [88] are also shown for comparison (after Ref. 76). 1698 Low Temperature Physics/Fizika Nizkikh Temperatur, 2017, v. 43, No. 12 Isotopic and spin-nuclear effects in solid hydrogens trum (Fig. 16), and from the condition / = 0∂ ∂η the or- der parameter was found as a function of temperature. Let us consider the situation for 0< <mP P P where or- dering at zero temperature is absent. As one can see from Fig. 18, there is a range of pressures where the ordered phase emerges with increasing temperature. Within this range of pressures the system undergoes two first-order phase transitions with increasing temperature, viz., the low-tem- perature one from disordered to ordered state and the high- temperature one from ordered to disordered state. The for- mer is of a quantum nature, QOM, the latter is the common orientational disordering. This behavior has the following explanation. At = 0T the only ordering factor is the poten- tial energy; as temperature increases the singlet of the splitted triplet start to get occupied increasing the contribu- tion of the entropy term (T S∆ ) of the free energy, when the sum of the contribution coming from the potential energy and the term T S∆ became bigger than the contribution from the kinetic energy a phase transition to the ordered Fig. 16. Energy spectrum of a hindered linear rotor in the field 2(cos ( ))VP ϑ . E and V are given in units of the rotational con- stant rotB . The numbers at the curves indicate the degeneracy of the even- and odd-J energy levels for H2, D2, and HD, respec- tively, taking into account nuclear spin degeneracy (after Refs. 80, 81, 89). Fig. 17. Temperature dependence of the entropy for free rotor and for the orientational ordered state of an all-J system in the field 2(cos( ))U P− η ϑ at different values of U (dot-dashed lines). The solid line in the inset shows the entropy difference between the ordered and disordered phases for / 12.9U B ≈ (after Ref. 81). Fig. 18. Temperature dependence of the order parameter η of a system of all-J rotors. The dash curve is the line of equilibrium phase transitions, the dotted curve is the loci of absolute instabi- lity of the orientationally ordered state (after Ref. 81). Fig. 19. The –U T phase diagram for a system of interacting rotors under the field 2(cos( ))VP− θ where =V Uη (after Refs. 80, 81). Low Temperature Physics/Fizika Nizkikh Temperatur, 2017, v. 43, No. 12 1699 Yuri A. Freiman and Yanier Crespo phase takes place. When temperature grows further higher levels start to get occupied, as a consequence S∆ arrive to a maximum and begin to decrease till the entropy of the disordered state became grater than the ordered one and a second phase transition to the disordered phase occurs. As a result the phase diagram of this system became reentrant as shown in Fig. 19 where the U T− phase diagram for systems of interacting rotors is plotted [80,81]. 4.3.2. The reentrant behavior from numerical simulations The first attempt to go beyond mean field approach and study the reentrant behavior in solid HD by using standard numerical simulations was made in Ref. 84. In this work the Path Integral Monte Carlo technique [91] (PIMC) was employed using two main approximations: (i) the authors consider asymmetric rotors with centers fixed at sites of fcc lattice instead of hcp, and (ii) their anisotropic interac- tion potential was limited by the pure EQQ interaction: 50 5 0 224 3/2 ( ) 1( ) = , (4 ) nn nn nn R B R R  Γ   π   (16) where 0 nnR is the nearest neighbor distance at zero tempera- ture. Even when the EQQ interaction is the leading term of the pair interaction potential its selection is essential for getting the right transition pressure [85]. It results that the anisotropic interaction potential used in this work [84] was too soft [85] and as a consequence the minimum transition pressure mP obtained was small 10mP ≈ GPa when com- pared with the experimental value = 53mP GPa (see Fig. 20). Recently a more accurate approach was imple- mented by Crespo et al. in Ref. 85. In this paper, the au- thors model the HD solid, using the constant pressure PIMC (CP–PIMC) formalism. The HD solid was simulated as an assembly of rigid molecules, positioned on both fcc or hcp lattices and having rotational and translational de- grees of freedom. The pair interaction potential used was essentially that of Cui et al. [92] pair Cui iso ani < < = ( ) ( , , ) ,ij i j ij i j i j U U R U+ κ Ω Ω∑ ∑ R (17) where iso ( )ijU R is the isotropic component, ijR is the vec- tor connecting the center points of two molecules, = ( , )i i iΩ θ φ , is a vector containing the inclination and azimuthal angles ( , )i iθ φ in spherical coordinates and κ is a rescaling factor. Two sources for the anisotropic coeffi- cients pair aniU were considered: first Schaefer et al. [93], denoted as Schaefer CuiU ; second, Burton et al. [94], denoted by Burton CuiU . In order to determine the BSP line of HD it is necessary to define an appropriate order parameter to measure and monitor the angular order. Suppose we know the structure of the ordered phase; then it is possible to define N unit vectors iu corresponding to the orientation of molecule on all sites = 1, ,i N and define the order parameter: ( ) 2 2 , =1 =1 1= 3 cos ( ) 1 , 2 M N p i m i m i O MN   ⋅ −     ∑∑ n u (18) where M is the total number of Trotter slices [91], ,i mn labels the orientation of the molecule i in the Trotter slice m for a given configuration visited with the PIMC algo- rithm. This order parameter measures the extent of order- ing (and by difference, its deviation) relative to a given orientational structure defined by the set { }iu and have been extensively used to study the BSP of solid hydrogens [84,85,92,95,96]. Nevertheless the order parameter in Eq. (18) suffers from an important limitation, namely: if the system reaches an ordered structure different from the reference one associated to the chosen set { }iu , still 0pO ≈ . Thus, the condition = 0pO is not sufficient to ensure that the system is in an orientationally disordered phase. To address this issue further, a second order parameter defined as 3 2Total =1 , =1 1 1= , 6 N Q jk i j k O N    ∑ ∑  (19) can be monitored, where ( ) 2 Total 2 , , 0 =1 1= 3 T MMC i i jk j a k a jk MC a r r r T M − δ∑ (20) with MCT is the MC time. Total jk is the quadrupolar moment of a system of MCT M molecules with a charge per atom of 1( )MCT M − [85]. If the molecules rotate in MC or Trotter time [91] (showing spherical symmetry), then Total = 0jk ; if, on the contrary, the molecules are frozen in a quadru- polar configuration both in MC and Trotter time, then 23 Total , =1 1 = 1 6 jkj k    ∑  . Thus QO〈 〉 signals quadrupolar Fig. 20. The temperature-pressure phase diagram of solid HD. The open circles are the experimental data [78]. The up open triangles are the constant volume PIMC results reported by Shin et al. [84]. All other symbols are constant pressure PIMC results obtained by Crespo et al. [85]. 1700 Low Temperature Physics/Fizika Nizkikh Temperatur, 2017, v. 43, No. 12 Isotopic and spin-nuclear effects in solid hydrogens order in a more general way than pO〈 〉 , which is restricted to a particular choice of { }iu . The inset of Fig. 20 shows the order parameter pO as a function of MC time at four pressures: 52 GPa (open circles), 54 GPa (open squares), 56 GPa (solid diamond) and 58 GPa (solid up triangles). The abrupt change seen in the order parameter indicates the presence of a phase tran- sition from a disordered phase to an ordered one, suggesting that HD, at a temperature of = 30 KT , orders at 58≈ GPa. Similar results were reported for the second order parame- ter Total jk [85]. Figure 20 shows the calculated tempera- ture-pressure phase diagram of solid HD, along with the experimental transition line (open circles, from Ref. 78), also compared to theoretical phase diagrams obtained in previous calculations. The up open triangles are the con- stant volume PIMC results reported by Shin et al. [84]. The rest of the symbols are CP-PIMC results considering different lattices (fcc, hcp) and different interaction poten- tial ( Schaefer CuiU and Burton CuiU ) obtained by Crespo et al. [85]. As can be seen the reentrant behavior for the HD phase line is obtained on both the realistic hcp and the fictitious fcc lattices, and for both choices of the anisotropic poten- tials. The minimum pressure at which the transition occurs is 56mP ≈ GPa for the hcp lattice with the Burton CuiU poten- tial, close to the experimental value of = 53mP GPa. Using alternatively Schaefer CuiU yields instead 70mP ≈ GPa. The edge point temperature = 25 KmT is also in good agree- ment with the experimental value of = 30 KmT . In gen- eral, the obtained BSP line with the Burton CuiU potential is in good agreement with the experiment. These results show the crucial role played by the choice of the EQQ interac- tion in determining the BSP transition pressure. Interest- ingly, the BSP transition line for the fcc lattice is strongly shifted downwards by about 30 GPa, regardless of the po- tential used. In that case the BSP structure is Pa3, and the stronger tendency to order is evidently due to the lack of frustration in this structure. In other words, it is the angular frustration present in the hcp lattice, and the connected poorer relative stability of the 2 /C c structure [85], that renders the BSP angular structure much more prone to melting than the 3Pa . To estimate the jump in entropy during this transition the slope of the BSP line can be used. In fact according to the Clapeyron equation / = /dP dT S V∆ ∆ , therefore di- rectly connected to the entropy jump S∆ between the low temperature quantum rotationally melted phase, and the higher temperature BSP solid phase. During this first order phase transition, the entropy jump vanish both at = 0T (because of the Nernst theorem), and at the reentrant edge point, where the entropies of both phases have the same value. In between, the entropy jump is finite. From the slope of the calculated phase line a maximum value of 0.4 0.2BS k∆ ± near 60 GPa at 15 K can be obtained, to be compared with 0.5 Bk of the experimental slope at same pressure and temperature (see Fig. 20) and the contribution to the entropy, predicted by mean field theory, in the strong interaction limit [81], equal to ln 2 = 0.693B Bk k , confirming that the entropy is an additional factor that sta- bilizes the ordered phase. 4.4. Reentrant phase transitions in ortho-para mixtures As was shown in the preceding section, depending on the parity of the rotational quantum number J, solid hyd- rogens exhibit either pressure-driven BSP quantum phase transitions (even-J species p-H2 and o-D2) or usual order- disorder classical phase transitions (odd-J species o-H2 and p-D2). Solid HD was found to possess an additional anom- aly: its BSP phase transition line displays a minimum, indi- cating that the disordered phase is reentrant. Such a strong difference in the pressure behavior of even-, odd-, and all-J species raises an intriguing question of the possible phase diagrams of their mixtures. Two theoretical papers were devoted to this problem. We will start from a mean-field theory developed by Freiman, Tretyak, Mao and Hemley [89]. To circumvent the main difficulty which one encounters considering mix- tures, the systems without a translational invariance, the authors used the following trick. It was supposed that each site of the lattice is occupied by a superposition of even-J and odd-J rotors. Two limiting cases are considered: mix- tures at thermodynamic equilibrium, where the conversion time is small or comparable with the thermalization time, and the opposite case, frozen mixtures, when the conver- sion time is large compared with all other relevant times. In the former case, the fractions of even and odd species in the mixture are temperature dependent and equal to ther- modynamically average concentrations; for the latter, the fraction of the species are fixed. The molecular field Ham- iltonian (15) was used to describe dynamics of orient- ational degrees of freedom. In the case of the frozen systems the partition function of the system is a product of the partition functions of the even and odd systems: even odd=Z Z Z . (21) In this case the free energy of the system is a sum of free energies of the even and odd systems, and there is a simple relation between the order parameter of the system 20= Yη 〈 〉 and order parameters of even even 20 even= Yη 〈 〉 and odd odd 20 odd= Yη 〈 〉 systems: 20 20 even 20 odd= (1 ) .Y c Y c Y〈 〉 − 〈 〉 + 〈 〉 (22) In the case of the equilibrium systems, the partition function of the system is the sum of the partition functions of even and odd systems: even odd= .Z Z Z+ (23) Low Temperature Physics/Fizika Nizkikh Temperatur, 2017, v. 43, No. 12 1701 Yuri A. Freiman and Yanier Crespo In this case a simple expression relating 20= Yη 〈 〉 with even 20 even= Yη 〈 〉 and odd odd 20 odd= Yη 〈 〉 does not exist. The resulting phase diagrams are shown in Figs. 21 and 22 for the thermodynamically equilibrium and frozen even-J odd-J mixtures, respectively. In Fig. 21, phase diagrams for the pure even-J and pure odd-J systems, as well as that for the all-J system, are also shown for comparison. As one can see, there is a striking difference between these diagrams. The equilibrium even-J odd-J mixtures exhibit the same type of reentrant behavior as known for the all-J systems (see Subsecs. 4.2, 4.3). The strongest reentrance is displayed by the even-J odd-J linear rotor mixture (equilibrium o-p-H2 mixture), the least one by the equilibrium o-p mixture of D2 with HD intermediate be- tween them. The frozen systems display qualitatively dif- ferent behavior of their phase transition lines (Fig. 21). For any even-J odd-J compositions excluding the case of the pure even-J system the phase transition lines go monoton- ically to the point = 0P with decreasing temperature. This means that the passage to the limit of the zero fraction of the odd modification is discontinuous. An analysis of the entropy contribution to the free ener- gy (Figs. 23, 24) furnishes an understanding of the nature of the reentrant behavior and of the distinctions between Fig. 23. Entropy of the equilibrium o-p mixture of solid hydro- gens (after Freiman et al. [89]). Fig. 22. Frozen o-p phase diagram (after Freiman et al. [89]). Fig. 24. Entropy of the frozen o-p mixture of solid hydrogens (after Freiman et al. [89]). Fig. 21. Equilibrium o-p phase diagram (after Freiman et al. [89]). 1702 Low Temperature Physics/Fizika Nizkikh Temperatur, 2017, v. 43, No. 12 Isotopic and spin-nuclear effects in solid hydrogens the equilibrium and frozen ortho-para mixtures. As can be seen from Fig. 23, the entropy of the equilibrium and all-J systems is anomalous: in the low-temperature region the entropy of the ordered phase is higher than that of the dis- ordered phase. Thus, the entropy contribution is an addi- tional ordering factor; its value increases with temperature and, as a result, the ordering pressure goes down as tem- perature increases. At a certain temperature, the entropy of the disordered phase becomes higher than that of the or- dered phase and the entropy contribution turns into a dis- ordering factor and the ordering pressure turns upward after passing the minimum point. In contrast with that, for the frozen systems no such anomaly exists for the frozen systems (see Fig. 25). In its turn, the entropy anomaly can be traced at the mi- croscopic level to characteristic features of the spectrum of the ordered phase (Fig. 16). For the all-J system (HD) the ground state is formed by two closely spaced levels (a doublet in the strong interaction limit) as opposite to the singlet ground state of the disordered phase. As a result, there is an additional contribution ln 2R to the ordered phase entropy. At the same time, both even- and odd-J pure systems have singlet ground states and their transition lines do not exhibit reentrance. The same entropy-based considerations can be applied to the case of the BSP transi- tions in the even-J odd-J mixtures of quantum linear ro- tors. In the equilibrium case the partition function of the homonuclear molecules is the sum of the partition func- tions of the even-J and odd-J states taken with their nu- clear spin- symmetry weights. As a result, for equilibrium mixtures the ground state phase of the ordered phase is a quadruplet for H2 (in a large V limit) and a octuplet for D2. Taking into account that the disordered phase entropies (at zero temperature) are zero for H2 and ln 4R for D2, additional contributions to the ordered phase entropies for H2 and D2 are ln 4R and ln(4 / 3)R , respectively. This explains why the reentrance decreases in the sequence H2, HD, D2. In the case of the frozen mixtures, the partition function of the mixture is a product of the partition func- tions of the components, each having a singlet ground state, with no reentrance resulting. Hetényi et al. [97], using a multi-parametrized mean- field (MF) formalism, also found that the equilibrium even-J odd-J mixtures have reentrant phase diagrams. The main advantage of their formalism is that contrary to the previous work where it was supposed that each site of the lattice is occupied by a superposition of even-J and odd-J rotors the new formalism allows for onsite distinc- tion of different spin-nuclear species that can be distributed randomly inside the lattice. This is a more realistic approx- imation that allows also for the inclusion of local order parameter correlation functions and therefore it is possible to differentiate short-range order from long-range order a feature that is absent in the standard MF theory where only phases of complete order or disorder are possible. The multiorder parameter (MOP) mean-field theory is based on the trial Hamiltonian 5 2 0 0 rot 20 =1 < 1ˆ= ( ) 2 N N i i i iji i j R H B L K Y R    + Ω − γ ×        ∑ ∑ * 40(224;00) ( ),j ijC Y× γ Ω (24) where K is the coupling strength, (224;00)C is a Clebsch– Gordan coefficient, ijΩ denote the direction of the vector connecting rotors i and j and iγ are parameters. Variation of the free energy leads to the self-consistent expression 20 0= ( )i iYγ 〈 Ω 〉 . (25) Fig. 25. Phase diagrams of the pure systems H2 and D2 at thermal equilibrium distribution calculated using multiorder parameter mean-field theory and the standard mean-field theory (MF). After Ref. 97. Fig. 26. Multiorder parameter MF phase diagrams for coupled quadrupolar rotor models corresponding to solid molecular hyd- rogen frozen at various ortho concentrations. After Ref. 97. Low Temperature Physics/Fizika Nizkikh Temperatur, 2017, v. 43, No. 12 1703 Yuri A. Freiman and Yanier Crespo Since the trial Hamiltonian in Eq. (24) is a sum of single- rotor Hamiltonians, it follows that for each i in Eq. (25) the average needs to be performed over the corresponding coordinate i only. Ortho-para distinction can be imple- mented by restricting a particular average to be over odd-J or even-J states. This formalism was applied to a system of coupled quadrupolar rotors whose centers of mass form a fcc lattice. The methodology for the self-consistent calcu- lation of the parameters { iγ } is described in Ref. 97. Order is then signaled by nonzero solutions for iγ , disorder by solutions in which all iγ are identically zero. The orientational ordering of the system with ortho-para concentrations at thermal equilibrium is shown in Fig. 26 along with the phase diagrams of the pure systems. For H2 and D2 the calculated phase diagrams via the MOP mean- field theory and via the standard MF theory are very simi- lar (see Fig. 21) and the reentrance phase line for H2 and D2 is also obtained within this approximation. In Fig. 26 the results of the formalism presented above are shown for solid H2 at a frozen concentration of ortho species. As in the case of standard MF as the ortho concentration is de- creased the system tends towards disorder entering the ordered state at higher coupling constants for a given tem- perature (see also Fig. 22). Nevertheless a noteworthy re- sult is obtained with this new formalism. As one can see from Fig. 26 even at an ortho concentration of 1% the sys- tem enters an ordered state at coupling constants quite dif- ferent from that of pure parahydrogen, and that for any ortho concentration the ground state is always ordered. This result is not observed in the case of standard MF theo- ry (see the 0.01 curve in Fig. 22). The reason for this dif- ference is because the orientational order is short ranged at low temperatures. This feature can be just capture with the MOP mean-field approximation where it is possible to dif- ferentiate short-range order from long-range order while only long-range order could be observed in standard MF theory. In order to assess the nature of the ordering the cor- relation functions ( )G r of the local order parameters { }iγ was calculated using the expression. 2( ) = [ (0) ( )]G r rγ γ . (26) The correlation functions for different temperatures along the reentrant phase diagram (Fig. 25) are shown for H2 in Fig. 27. As the temperature increases correlation increases along the phase boundary. At high temperatures ( / > 1T B ) the order is definitely long range. The short-range order is expected to be present at low-T up to / < 0.75T B . The increases of the correlation along the phase boundary with increasing temperature indicates the onset of long- range order. The onset of long-range order is due to the fact that ortho-para distinction ceases as temperature and coupling constant (pressure) are increased. 5. Conclusion In this Section we summarize the results discussed in this review article with a stress on open experimental and theoretical problems that still remain to be solved. These problems span the whole region of the phase diagram. Naturally, the most “hot” are problems at high-pressure frontiers. In the Introduction (Sec. 1) we have discussed the con- straints which quantum mechanics impose on the possible link between rotational states of the hydrogen molecules and their total nuclear spin. This symmetry-related link gives rise the existence of ortho- and paramodifications dis- playing large differences in the solid-state properties. In Sec. 2 we reviewed experimental and theoretical works devoted to studies of the ortho-para conversion in solid hydrogen and deuterium with the emphasis on the high-pressure studies. The highest pressures for which the experimental conversion rate data exist are at present 58 GPa. The extension of these studies to higher pressures would provide a critical test of the conversion mechanisms proposed by theory. It will need further theoretical studies to validate the conception of ortho-para states at ultra-high pressures. In Sec. 3 we discuss two of numerous effects displaying by the = 0 = 1J J− mixtures of solid hydrogen and deuter- ium: the effect of = 1J admixtures on the molar volume, and the hcp lattice distortion. Hydrogen is a quantum solid and is therefore highly compressible. Due to the large compressibility, the weak quadrupole-quadrupole forces related with = 1J species have a non-negligible effect on the molar volume. As compared with fcc, the hcp lattice has an additional degree of freedom associated with the /c a ratio. In this Section we discuss the deviation of the axial ratio for the = 0 = 1J J− mixtures from the ideal value 8 / 3 1.633≈ . It was shown that the presence of = 1J molecules increas- es the hcp lattice distortion by two order of magnitude. Fig. 27. Correlation functions along the phase boundary of the so- lid H2 at thermal equilibrium. After Ref. 97. 1704 Low Temperature Physics/Fizika Nizkikh Temperatur, 2017, v. 43, No. 12 Isotopic and spin-nuclear effects in solid hydrogens Section 4 is devoted to the peculiarities of the phase di- agram of solid HD. Near 50 GPa it exhibits a reentrant phase transition where a rotationally ordered («broken sym- metry») crystalline phase surprisingly transforms into a orientationally disordered high-symmetry phase upon cool- ing. The molecular-field theory gives the qualitative reason for reentrance: the higher entropy of the broken symmetry phase, due to the inequivalence of H and D, as opposed to the low entropy of the high-symmetry phase where the rotational melting is quantum mechanical — a Pome- ranchuk-like mechanism. The entropy jump across the tran- sition is found to be comparable with ln 2, the value expect- ed for the Pomeranchuk mechanism. Path Integral Monte Carlo calculations give a reentrant phase boundary in good agreement with experiment. On the ground of the similar entropy-based considera- tions the molecular-field theory forecasts that the broken- phase transitions in thermodynamically equilibrium ortho- para mixtures of homonuclear species of H2 and D2 should exhibit reentrant phase transition lines. This prediction makes the respective experimental studies and the devel- opment of theory of the BSP transition in such systems beyond the MF approximation important actual problem In Sec. 5 we listed a number of problems both in exper- iment and in theory that still remain to be solved. These problems span the whole region of the phase diagram. Naturally, the most “hot” are problems at the high-pressure frontiers. The authors are deeply indebted to Erio Tosatti and Mi- khail Strzhemechny for numerous fruitful discussions. eferences 1. I.F. Silvera, Rev. Mod. Phys. 52, 393 (1980). 2. J. Van Kranendonk, Solid Hydrogen, Plenum, New York (1983). 3. Physics of Cryocrystals, V.G. Manzhelii and Yu.A. Freiman (eds.), AIP Press, New York (1997). 4. E. Cremer and M. Polanyi, Trans. Faraday Soc. 28, 435 (1932); Z. Phys. Chem. B 21, 459 (1933). 5. L. Farkas, Ergeb. Exacten Naturwiss. 12, 163 (1933). 6. A. Driessen, E. van der Poll, and I.F. Silvera, Phys. Rev. B 30, 2517 (1984). 7. G. Ahlers, J. Chem. Phys. 40, 3123 (1964). 8. P. Pedroni, H. Meyer, F. Weinhaus, and D. Haase, Solid State Commun. 14, 279 (1974). 9. F. Schmidt, Phys. Rev. B 10, 4480 (1974). 10. R.F. Buzerak, Ph.D. Thesis, Duke University (1975). 11. R. Jochemsen, Ph.D. thesis, Univ. of Amsterdam (1978). 12. I.F. Silvera, P.J. Berkhout, and L.M. van Aernsbergen, J. Low Temp. Phys. 35, 611 (1979). 13. R.J. Hemley, A.F. Goncharov, H.-k. Mao, J.H. Eggert, and E. Karmon, J. Low Temp. Phys. 110, 75 (1998). 14. J.H. Eggert, E. Karmon, R.J. Hemley, H.-k. Mao, and A.F. Goncharov, Proc. Natl. Acad. Sci. USA 96, 269 (1999). 15. M. Pravica, and I.F. Silvera, Phys. Rev. Lett. 81, 4180 (1998). 16. M.A. Strzhemechny and R.J. Hemley, Phys. Rev. Lett. 85, 5595 (2000). 17. K. Motizuki and T. Nagamiya, J. Phys. Soc. Jpn. 11, 93 (1956). 18. A.J. Berlinsky and W.N. Hardy, Phys. Rev. B 8, 5013 (1973). 19. A.J. Berlinsky, Phys. Rev. B 12, 1482 (1975). 20. M.A. Strzhemechny, R.J. Hemley, H.-k. Mao, A.F. Goncharov, and J.H. Eggert, Phys. Rev. B 66, 014103 (2002). 21. K. Motizuki, J. Phys. Soc. Jpn. 12, 163 (1957). 22. W.N. Hardy and A.J. Berlinsky, Phys. Rev. B 8, 4996 (1973). 23. Yu.Ya. Milenko and R.M. Sibileva, Fiz. Nizk. Temp. 1, 796 (1975) [Sov. J. Low Temp. Phys. 1, 382 (1975)]. 24. P.J. Berkhout, J. Th. Minneboo, and I.F. Silvera J. Low Temp. Phys. 32, 401 (1978). 25. M. Calkins, R. Banke, X. Li, and H. Meyer, J. Low Temp. Phys. 65, 47 (1986). 26. M.I. Bagatskii, A.I. Krivchikov, V.G. Manzhelii, I.Ya. Minchina, and P.I. Muromtsev, Fiz. Nizk. Temp. 13, 1001 (1987) [Sov. J. Low Temp. Phys. 13, 571 (1987)]. 27. M.A. Strzhemechny and O.I. Tokar, Fiz. Nizk. Temp. 11, 813 (1985) [Sov. J. Low Temp. Phys. 11, 446 (1985)]. 28. M.A. Strzhemechny and R.J. Hemley, Fiz. Nizk. Temp. 29, 941 (2003) [Low Temp. Phys. 29, 703 (2003)]. 29. I.F. Silvera, J. Low Temp. Phys. 112, 237 (1998). 30. I.F. Silvera and M. Pravica, J. Phys.: Condens. Matter 10, 11169 (1998). 31. Y. Akahama, M. Nishimura, H. Kawamura, N. Hirao, Y. Ohishi, and K. Takemura, Phys. Rev. B 82, 060101(R) (2010). 32. N.W. Ashcroft, J. Non-Cryst. Solids 156, 621 (1993). 33. I.F. Silvera, W.N. Hardy, and J.P. McTague, Phys. Rev. B 8, 1578 (1972). 34. A. Driessen, J.A. de Waal, and I.F. Silvera, J. Low Temp. Phys. 34, 255 (1979). 35. R. Howard, Phys. Lett. A 32, 37 (1970). 36. P. Schwerdtfeger, N. Gaston, R.P. Krawczyk, R. Tonner, and G.E. Moyano, Phys. Rev. B 73, 064112 (2006). 37. F.H. Stillinger, J. Chem. Phys. 115, 5208 (2001). 38. Yu.A. Freiman, S.M. Tretyak, A. Grechnev, A.F. Goncharov, J.S. Tse, D. Errandonea, H.-k. Mao, and R.J. Hemley, Phys. Rev. B 80, 094112 (2009). 39. Yu.A. Freiman, S.M. Tretyak, F. Jezowski, and R.J. Hemley, J. Low Temp. Phys. 126, 703 (2002). 40. Yu.A. Freiman, Alexei Grechnev, S.M. Tretyak, Alexander Goncharov, and Russell J. Hemley, Phys. Rev. B 86, 014111 (2012). 41. A. Grechnev, S.M. Tretyak, and Yu.A. Freiman, Fiz. Nizk. Temp. 36, 423 (2010) [Low Temp. Phys. 36, 333 (2010)]. 42. W.H. Keesom, J. de Smedt, and H.H. Moon, Leiden Com- mun. 19, Suppl. No. 69, No. 209d (1929/1930). 43. I.N. Krupskii, A.I. Prokhvatilov, and G.N. Shcherbakov, Fiz. Nizk. Temp. 9, 83 (1983) [Sov. J. Low Temp. Phys. 9, 42 (1983)]; ibid. 10, 5 (1984) [ibid. 10, 1 (1984)]. 44. S.N. Ishmaev, I.P. Sadikov, A.A. Chernyshov, B.A. Vindryaevskii, V.A. Sukhoparov, A. Telepnev, and G.V. Kobelev, Sov. Phys. JETP 57, 228 (1983). 45. S.N. Ishmaev, I.P. Sadikov, A.A. Chernyshov, B.A. Vindryaevskii, V.A. Sukhoparov, A. Telepnev, G.V. Kobelev, and R.A. Sadikov, Sov. Phys. JETP 62, 721 (1985). Low Temperature Physics/Fizika Nizkikh Temperatur, 2017, v. 43, No. 12 1705 Yuri A. Freiman and Yanier Crespo 46. A.F. Goncharov, M.A. Strzhemechny, H.-k. Mao, and R.J. Hemley, Phys. Rev. B 63, 064304 (2001). 47. Yu.A. Freiman, S.M. Tretyak, Alexei Grechnev, Alexander F. Goncharov, and Russell J. Hemley, Phys. Rev. B 90, 024501 (2014). 48. P. Loubeyre, R. LeToullec, D. Hausermann, M. Hanfland, R.J. Hemley, H.-k. Mao, and L.W. Finger, Nature 383, 702 (1996). 49. R.J. Hemley, H.-k. Mao, L.W. Finger, A.P. Jephcoat, R.M. Hazen, and C.S. Zha, Phys. Rev. B 42, 6458 (1990). 50. H.-k. Mao and Russell J. Hemley, Rev. Mod. Phys. 66, 671 (1994). 51. H. Kawamura, Y. Akahama, S. Umemoto, K. Takemura, Y. Ohishi, and O. Shimomura, J. Phys.: Condens. Matter 14, 10407 (2002); ibid. 86, 014111 (2012). 52. Yu.A. Freiman, S.M. Tretyak, Alexander F. Goncharov, H.-k. Mao, and Russell J. Hemley, Fiz. Nizk. Temp. 37, 1302 (2011) [Low Temp. Phys. 37, 1038 (2011)]. 53. K. Mendelssohn, M. Ruhemann, and F. Simon, Z. Phys Chem. B 15, 121 (1931). 54. K. Schaefer, Z. Phys. Chem. B 42, 380 (1939); ibid. B 45, 451 (1940). 55. T. Nagamiya and H. Kisi, Busseiron Kenkyu (in Japanese) 39, 64 (1951). 56. T. Nakamura, Prog. Theor. Phys. (Kyoto) 14, 135 (1955). 57. R.W. Hill and B.W. Richetson, Philos. Mag. 45, 277 (1954). 58. R.J. Roberts, E. Rojas, and J.G. Daunt, J. Low Temp. Phys. 24, 265 (1976). 59. W.H. Orttung, J. Chem. Phys. 36, 652 (1962). 60. J. van Kranendonk and V.F. Sears, Can. J. Phys. 44, 313 (1966). 61. A.B. Harris and E.R. Hunt, Phys. Rev. Lett. 16, 845 (1966). 62. W.N. Hardy and J.R. Gaines, Phys. Rev. Lett. 19, 1417 (1967). 63. J.H. Constable and J.R. Gaines, Solid State Commun. 9, 155 (1977). 64. R. Schweizer, S. Washburn, and H. Meyer, J. Low Temp. Phys. 37, 289 (1979). 65. W.N. Hardy, A.J. Berlinsky, and A.B. Harris, Can. J. Phys. 55, 1150 (1977). 66. J.C. Raich and L.B. Kanney, J. Low Temp. Phys. 28, 95 (1977). 67. S. Luryi and J. van Kranendonk, Can. J. Phys. 57, 933 (1979). 68. R.M. Dickinson, T.J. Byers, and T. Oka, J. Low Temp. Phys. 102, 241 (1996). 69. A.I. Prokhvatilov, M.A. Strzhemechny, and G.N. Shcherbakov, Fiz. Nizk. Temp. 19, 622 (1993) [Low Temp. Phys. 19, 445 (1993)]. 70. I.N. Krupskii, A.I. Prokhvatilov, and G.N. Shcherbakov, Fiz. Nizk. Temp. 9, 858 (1983) [Sov. J. Low Temp. Phys. 9, 446 (1983)]. 71. M.A. Strzhemechny, Phys. Rev. B 73, 174301 (2006). 72. J. Bigeleisen, J. Chem. Phys. 39, 769 (1963). 73. M. Trefler and H.P. Gush, Phys. Rev. Lett. 20, 703 (1968). 74. M.A. Strzhemechny and R.J. Hemley, Phys. Rev. B 91, 144102 (2015). 75. Yu.A. Freiman and M.A. Strzhemechny, In: Physics of Cryo- crystals [3], Ch. 17, Kinetic Properties, AIP Press, NY (1997), p. 489. 76. A. Chijioke and I.F. Silvera, Phys. Rev. Lett. 97, 255701 (2006). 77. L. Cui, N.H. Chen, and I.F. Silvera, Phys. Rev. B 51, 14987 (1995). 78. F. Moshary, N.H. Chen, and I.F. Silvera, Phys. Rev. Lett. 71, 3814 (1993). 79. H.E. Lorenzana, I.F. Silvera, and K.A. Goettel, Phys. Rev. Lett. 64, 1939 (1990). 80. Yu.A. Freiman, V.V. Sumarokov, A.P. Brodyanskii, and A. Jezowski, J. Phys.: Condens. Matter 3, 3855 (1991). 81. A.P. Brodyanskii, V.V. Sumarokov, Yu.A. Freiman, and A. Jezowski, Fiz. Nizk. Temp. 19, 520 (1993) [Low Temp. Phys. 19, 368 (1993)]. 82. Yu.A. Freiman, and S.M. Tretyak, A. Jezowski, and R.J. Hemley, J. Low Temp. Phys. 113, 723 (1998). 83. Yu.A. Freiman, S.M. Tretyak, and Andrzej Jezowski, J. Low Temp. Phys. 110, 147 (1998). 84. Hyeondeok Shin and Yongkyung Kwon, J. Korean Phys. Soc. 54, 1582 (2009). 85. Yanier. Crespo, Alessandro Laio, Giuseppe E. Santoro, and Erio Tosatti, Phys. Rev. B 84, 144119 (2011). 86. R.J. Hemley and H.-k. Mao, Phys. Rev. Lett. 61, 857 (1988). 87. R.J. Hemley and H.-k. Mao, Phys. Rev. Lett. 63, 1393 (1989). 88. L. Cui, N.H. Chen, S.J. Jeon, and I.F. Silvera, Phys. Rev. Lett. 72, 3048 (1994). 89. Yu.A. Freiman, S.M. Tretyak, H.-k. Mao, and R.J. Hemley, J. Low Temp. Phys. 139, 765 (2005). 90. V.V. Sumarokov and Yu.A. Freiman, Fiz. Nizk. Temp. 13, 197 (1987) [Sov. J. Low Temp. Phys. 13, 109 (1987)]. 91. D.M. Ceperley, Rev. Mod. Phys. 67, 279 (1995). 92. T. Cui, E. Cheng, B.J. Alder, and K.B. Whaley, Phys. Rev. B 55, 12253 (1997). 93. J. Schaefer and E. Köhler, Z. Phys. B 13, 217 (1989). 94. P.G. Burton and U.E. Senff, J. Chem. Phys. 76, 6073 (1982). 95. K.J. Runge, M.P. Surh, C. Mailhiot, and E.L. Pollock, Phys. Rev. Lett. 69, 3527 (1992). 96. M.P. Surh, K.J. Runge, T.W. Barbee, E.L. Pollock, and C. Mailhiot, Phys. Rev. B 55, 11330 (1997). 97. B. Hetényi, S. Scandolo, and E. Tosatti, Phys. Rev. Lett. 94, 125503 (2005). 1706 Low Temperature Physics/Fizika Nizkikh Temperatur, 2017, v. 43, No. 12 1. Introduction 2. Ortho-para conversion 2.1. Experiment 2.2. Theoretical considerations 2.3. Conversion in solid D2 2.4. Validity of the concept of ortho-para states at high pressure 3. Thermodynamic properties of J = 0 – J = 1 mixtures 3.1. Molar volume 3.2. HCP lattice distortion in solid H2 as a function of o-p composition 3.3. Experimental determination of the crystal field splitting of isolated impurities in solid parahydrogen 3.4. HCP lattice distortion in solid HD 4. Broken symmetry phase transition in solid HD and o-p mixtures 4.1. Introduction 4.2. Experimental results 4.3. Theoretical studies 4.3.1. Quantum orientational melting in all-J system 4.3.2. The reentrant behavior from numerical simulations 4.4. Reentrant phase transitions in ortho-para mixtures 5. Conclusion eferences