Next Article in Journal
Revisiting Alloy Design of Al-Base Alloys for Potential Orthotics and Prosthetics Applications
Next Article in Special Issue
Recent Progress in Phase Stability and Elastic Anomalies of Group VB Transition Metals
Previous Article in Journal
Effect of Process Parameters on the Microstructure of Aluminum Alloys Made via Ultrasonic Additive Manufacturing
Previous Article in Special Issue
Effects of Physical and Chemical Pressure on Charge Density Wave Transitions in LaAg1−xAuxSb2 Single Crystals
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Corresponding States for Volumes of Elemental Solids at Their Pressures of Polymorphic Transformations

Department of Geoscience, University of Nevada Las Vegas, Las Vegas, NV 89154, USA
Crystals 2022, 12(12), 1698; https://doi.org/10.3390/cryst12121698
Submission received: 1 November 2022 / Revised: 15 November 2022 / Accepted: 16 November 2022 / Published: 23 November 2022
(This article belongs to the Special Issue Pressure-Induced Phase Transformations (Volume II))

Abstract

:
Many non-molecular elemental solids exhibit common features in their structures over the range of 0 to 0.5 TPa that have been correlated with equivalent valence electron configurations. Here, it is shown that the pressures and volumes at polymorphic transitions obey corresponding states given by a single, empirical universal step-function Vtr/L = −0.0208(3) · Ptr + Ni, where Vtr is the atomic volume in Å3 at a given transformation pressure Ptr in GPa, and L is the principal quantum number. Ni assumes discrete values of approximately 20, 30, 40, etc. times the cube of the Bohr radius, thus separating all 113 examined polymorphic elements into five discrete sets. The separation into these sets is not along L. Instead, strongly contractive polymorphic transformations of a given elemental solid involve changes to different sets. The rule of corresponding states allows for predicting atomic volumes of elemental polymorphs of hitherto unknown structures and the transitions from molecular into non-molecular phases such as for hydrogen. Though not an equation of state, this relation establishes a basic principle ruling over a vast range of simple and complex solid structures that confirms that effective single-electron-based calculations are good approximations for these materials and pressures The relation between transformation pressures and volumes paves the way to a quantitative assessment of the state of very dense matter intermediate between the terrestrial pressure regime and stellar matter.

1. Introduction

In difference to gases whose states can be described by the ideal gas- or the van der Waals equations of state, solids seem to defy the concept of a general equation of state. The range of structures and properties as well as their changes upon compression appear too large to be comprised within a single formula. Over a range of pressure of 0 to 500 GPa, many elements that are molecular solids at ambient conditions transform into atomic metals [1], while others that are normal metals at ambient conditions become semiconductors [2,3,4], and some monatomic elemental solids assume complex structures under compression that do not exist at ambient pressure [5,6].
These changes are not arbitrary but reflect general trends in the effect of pressure on the valence electron structure of solid matter over large pressure intervals [1,2,3,4,5,6,7,8,9,10,11,12]. The changes in the valence electron configuration reflect the different response of different orbital states to the increase in electron density upon compression [2,7,10,11,12,13,14,15]. The atomic volume often is markedly reduced upon such transitions. The transition from the bcc to the hcp structure in many transition metal elements is correlated with an increase in the hybridization of filled d-states with p-states of the next higher principal quantum number [16,17]. Similarly, the pressure-induced transformations of lanthanides from hcp to the Sm type to dhcp to fcc is explained through the increased occupancy of valence d-states [11,12,13], whereas the marked volume collapse and transitions to low symmetric phases in lanthanides are assigned to delocalization of the 4f states [12,13,15]. A different type of delocalization occurs in the low-Z elemental alkaline metals, which at high pressure transform into electrides in which inner shell electron states overlap with valence s-states and non-bonding p-states are occupied on expenses of the bonding s- and p-hybrid states [2,4,14]. As a consequence, elements such as Li and Na, which are normal metals at ambient conditions, become semiconductors at pressures between 100 and 200 GPa [2,3,4]. Changes in the electronic valence states have been studied quantitatively for some alkaline metals, transition group elements, and lanthanides through X-ray spectroscopy [10,14,15,16,17]. These changes are correlated with structural transformations. It has been noted that these pressure-induced valence changes follow general correlations of the nuclear charge number with the pressure, volume, and structure type [5,8,9,10]. In addition, there are correlations between the atomic volumes of equivalent polymorphs of different elements [8,18].
Here, it is shown that all known non-molecular elemental solids follow a rule of corresponding states if only the volumes and pressures at the polymorphic transitions are considered. This observation extends to high-pressure atomic polymorphs of elements that are molecular at ambient conditions, such as O, N, S, P, Cl, Br, and I. Only experimental data are considered here (see Section 2), hence the relation presented is without any kind of simplifying assumptions or theory.

2. Materials and Methods

Only published volumes and transformation pressures of polymorphs of elemental solids are used in this paper. Recently published data and work conducted under hydrostatic or nearly hydrostatic conditions are given preference. This includes data obtained from samples embedded in He or Ne as pressure media or data from crystals grown from melt at a high pressure. Data from compression experiments without pressure media are only included if the average pressure has been below 10 GPa. However, the induced transformations in Ir, Pb, and Th under non-hydrostatic stresses are included for reference. Similarly, compression data of some lanthanides are included due to their principal interest, although most of these data were not and possibly could not be acquired with the use of hydrostatic media. The plotted volumes Vtr and pressures Ptr at the polymorphic transitions are generally those of the first observation of a high-pressure polymorph upon compression at 300 K. In a few cases in which the transformations exhibits hystereses larger than 5 GPa, the arithmetic middle pressure has been taken as Ptr. In all other cases, the hysteresis has been added to the uncertainty. For some elemental high-pressure polymorphs, the structures have been redetermined in more recent studies but without reassessment of the 300 K isotherms. In such cases, the atomic volume of the correct structure has been used to reassess the atomic volume at the pressure of the transformation that has been reported in the earlier studies. The complete data are given in Table 1 at the end of the paper.

3. Results

For all of the examined 113 elemental solids, the volume at the phase transition Vtr = V(Ptr) (Ptr = transition pressure) divided by the principal quantum number L exhibits a universal linear pressure dependence of −0.0208(3) · Ptr (Ptr in GPa; atomic volume in Å3; the adjusted R2 of the linear fit was 0.9985; see Supplemental Figure S1). This is shown in Figure 1. Furthermore, within the uncertainties, the constant term of this equation provides a volume that for all of the examined elemental solids is nearly equal to n · 10 · rB3, where n is an integer number between 2 and 6, and rB is the Bohr radius in Å. This is shown in Figure 2. In other words:
Vtr/L + 0.0208(3) · Ptr = Ni ≈ n · 10 · rB3
The actual values of the five distinct volumes for Ni are 2.88(8), 4.58(4), 5.51(5), 7.14(5), 8.53(6) Å3/at, which is 19.5(4), 30.9(4), 37.2(8), 48.2(1), and 57.6(5) times the cube of the Bohr radius rB3 in Å3, respectively. Therefore, Figure 1 and Figure 2 show that the volumes of elemental solids at their polymorphic transitions assume corresponding states. It is important to recall that this universal relation of corresponding states only holds for the pressures and volumes at the phase transitions: Within the stability field of each polymorph, the pressure–volume relation is controlled by its specific compressibility that itself depends on pressure and temperature and that, in general, is far from being linear, universal, and discrete. Equation (1) is therefore not an equation of state.

4. Discussion

The ‘collective quantization’ shown in Figure 1 may, upon first glance, be assigned to the factorization by the principal quantum number L. However, the series of corresponding states Ni does not follow the rows of the periodic table. Rather, many elements with subsequent polymorphic transitions change between states Ni. For instance, the high-pressure polymorphs fcc, cl16, and oc120 of Na reside in N4 = 5.51(5) Å3/at ≈ 40 · rB3 along with K-IV, Rb-IV, -V, Cs-IV, and Ca-III to -IV; whereas the h4p-type Na resides in N5 = 7.14(5) Å3/at ≈ 50 · rB3 along with Rb-II and -III. In some cases, for instance Si, all polymorphs remain within one set Ni; however, for most elements at least some polymorphs fall into different sets. For instance, Cs-II and Cs-III are in N6 but Cs-IV and -V belong to N4 and Cs-VI to N3. A sequence of stepwise transitions to sets with a lower state Ni along with an increasing transformation pressure is found for many elements: the high-pressure polymorphs of Rb fall into the sets N5 (Rb-II and -III), N4 (Rb-IV, -V, -VIII), and N3 (Rb-VI, -VII); Li into N6 (Li-II), N4 (Li-III, -IV,-VI), and N3 (Li-V); and the various polymorphs of the alkaline earths Ca, Sr, and Ba are found in the sets N3 to N6 (Figure 1 and Figure 2), where the higher pressure polymorphs belong to sets with a smaller Ni. Lower sets Ni have markedly smaller volumes at a given pressure than the corresponding higher sets Ni+1 (Figure 1), and the polymorphic transitions that involve changes Ni → Ni−1 are accompanied by marked volume contractions (VLPtr − VHPtr)/V0 where VLPtr and VHPtr are the atomic volumes of the low- and the high-pressure polymorphs, respectively, at the pressure of the phase transitions; and V0 is the volume of the ambient pressure phase at standard conditions. A smaller number of elements also exhibit transitions into higher sets Ni+1 at higher pressure. In particular, this holds for the polymorphic transitions to Sc-V [19], bcc-Ti [20,21], hp4-Na [3], and Rb-VIII (Figure 1), but also for the elastic anomaly of Os around 290 GPa [22], which is taken here for an isostructural transition. There is a simple reason for these ‘upward transitions’ Ni → Ni+1: at the given transformation pressure, the corresponding volumes of the next lower set Ni−1 would be negative. It is to be expected that for all elements at sufficiently high pressures, the cascade of successive, markedly contractive transitions Ni → Ni−1 is followed by transitions Ni → Ni+1 to polymorphs in higher sets with a less pronounced contraction once Ni → Ni−1 reaches the limit of negative volumes. Transitions to higher sets come with smaller contractions (VLPtr − VHPtr)/V0 than those into lower sets.

4.1. Theoretical Explanation

‘Upward’ and ‘downward’ transitions between different states Ni can be correlated to changes in the valence electron configuration wherever such information has been obtained experimentally [10,15,17] or through calculation [14].The fact that the volumes Ni are nearly equal to integer multiples of rB3 has a straightforward theoretical basis: It implies that the volumes that occur at polymorphic transitions of non-molecular elements are controlled by the number of valence electrons rather than by their interactions. Besides the limited experimental accuracy the electron exchange- and Coulomb interactions account for part of the deviations from Equation (1) and from integer Ni values; thus, the interaction terms are of the second order with respect to the volumes. In other words, Equation (1) is the equivalent to an ideal equation of state for elemental solids, only that it does not describe continuous pressure–volume changes but discrete changes that occur at the phase transitions. This finding can be further quantified by considering the effect of the different valence orbital states as based on the experimentally determined pressure-induced hybridization of d-states of the alkaline metals Rb, Cs, and K (the latter element exhibits no measurable d-state occupancy up to 40 GPa) [10]. Figure 1 may be re-parametrized in terms of pressure-induced changes in the d-state occupancy at the valence level. In particular, the d-state occupancy of the elemental polymorphs can be correlated with the five sets N2 to N6 shown in Figure 1 through their 0 GPa intercepts. This is shown in Figure 3a. For instance, the d-state occupancy of Cs increases for phase Cs-II to -VI from ~0.0 to ~0.6, and for Rb from ~0.2 to 0.4 electrons between Rb-II and -VI [10]. The high-P polymorphs of Li and Na have little or no occupied d-states in this diagram and according to earlier computational work [2,4,14]. The correlation between Δn(d) and Ni is linear, at least for the polymorphs K-II, Rb-II,-IV, Cs-II, and Cs-III (black squares in Figure 3a): Within uncertainty, the slopes are equal for all sets. In terms of multiples of the electron elemental charge, this linear correlation between d-state occupancy, Vtr, and L is:
Ni = Ni0 − 6.11(14) · 10−11 · n(d) · e
where Ni0 = Ni at n(d) = 0, the constant factor is given in m3/(C · at), and e is the electron elemental charge. However, the correlation between Ni and n(d) is not continuous: Equations (1) and (2) can be combined to give:
V tr L = 2.08 ( 3 ) · 10 32 · P 6.11 ( 14 ) · 10 11 · i = 0 m 1 i · n ( d ) · e + N m
where Nm is the value assumed at n(d) = 0, P is in GPa, V is in m3/at, and e is the electron elemental charge. The effect of the d → p transfer on shifts into sets with a higher Ni has been discussed above. Overall, there are fewer data on pressure-induced changes in the valence p-state occupancy (summarized in Supplementary Figure S2, see refs. [10,16,17]), but within uncertainties, the relation between Vtr/L and n(p) has the same slope as in (3) but with an opposed sign (consistent with the transfer into higher sets Ni, whereas changes in the d-state occupancy generally result in changes to a lower Ni). Implementing n(p) into (3), dividing through rB3, and using the approximation Ni rB3 ~ 10⋅m gives:
V tr L · r B 3 = 0.1405 ( 20 ) · P + 10 · m 66.1 ( 15 ) · i = 0 m 1 i · 0.15 ( 4 ) + 66 ( 2 ) · j = p m j · 0.09 ( 21 ) = 0.1405 ( 20 ) · P + 10 · m 10.0 ( 6 ) · i = 0 m 1 i + 6 ( 1 ) · j = p m j
where the pressure is given in GPa, the volumes Ni are taken as integer multiples 10 · i ∈ IN (the set of natural numbers) of rB3 per atom, p ≤ n(d), and the increments Δn(d) are taken as 0.151; hence: 66.1 · 0.151 = 10 d-electrons and accordingly 66(2). 0.09(1) = 6(1) for the p-electrons. Thus, within uncertainties, the factors of the sum terms equal the number of p- and d-states and an additional term for f-electrons with factor −14 may be added. The actual values Ni may be substituted for 10m.
The volumes of elemental atomic solids at ambient conditions are not bound to Equation (1) because they do not represent volumes at phase transitions. However, the correlation V = Ni · L is within ±10% of the observed volumes at ambient conditions for 28 and within 20% for 38 out of the 42 non-molecular solid elements (Figure 3b). Considering that the atomic volumes of elemental metals are equal to the cube of the Wigner–Seitz radii, their approximate relation to Ni · L is equivalent to a relation recently found for the ionic radii [23]. However, in the present case of elemental solids the relation does not rely on any approximate model of bonding.

4.2. Molecular Gap and Forbidden Zone

There are two regions in Figure 1 that contain no values Vtr/L: (1) There is no observed Vtr/L equal to or smaller than 1.48 Å3/at at any pressure; 1.48 Å3/at equals 10 times the cube of the Bohr radius (per atom; this limiting value 10 rB3/at is defined here as N1). Upon compression, elemental solids such as Au and Re reach this limit (see Figure 4) but without further known transformations [24,25]. There is no known elemental solid that undergoes a polymorphic transition between 0 and 300 GPa where the volume at the transition is equal to or smaller than 10 rB3.
(2) The area between P ~ 20 GPa for Vtr/L ~ 8 Å3/at and 200–300 GPa for ~3 Å3/at contains no observations. This empty range has a nearly parabolic shape in Figure 1. It corresponds to volumes and pressures of molecular elemental solids (Figure 4).
For instance, the volumes Vtr of the molecular polymorphs β, γ, δ, and ε of oxygen as well as the metallic molecular phase ζ-O2 [26], are not compliant with Equation (1). The isotherm of ε-O2 intersects Vtr/L + N6 at 24 GPa without any observed transition ([26]; Figure 4). The isotherm of H2 intersects the Vtr/L + N6 at 26 GPa and Vtr/L + N5 at 54 and 135 GPa and again Vtr/L + N6 around 250 GPa, which is at the pressure of the transition to H2-IV ([27], Figure 4), but neither of these intersections corresponds to the transition to an atomic metal [27]. However, the transition from molecular N2 to the polymeric, semiconducting cg-phase of nitrogen [28] falls onto the N4 trajectory (Figure 1, Table 1). Cl2 intersects Vtr/L + N6 at 260 GPa, where it has been reported to become an atomic metal with an In-type structure [29] and thus follows the behavior predicted by Equation (1). Hence, the gap between 50–180 GPa for Vtr/L + N5 and N6 indirectly reflects the strength of the molecular bonds of these materials under compression and should be called a ‘molecular gap’. The intersection of the 300 K isotherm of the chlorine phase III (mc8; [29]) with Vtr/L + N5 at 180 GPa may be coincidental because this phase is probably still molecular and the Vtr of the intermediate phase IV (i-oF3, [29]) falls in between Vtr/L + N5 and Vtr/L + N6. The isotherms of N, S, P, Br, and I all intersect the trajectories Vtr/L + Ni at the pressures of the observed transitions to non-molecular phases [8,27,30,31] and fall mostly below the ‘molecular gap’ because their transitions to non-molecular solids occur at lower pressure.
Interestingly, the extrapolated pressure of the direct gap closure of H2 that has been assessed through inelastic X-ray spectroscopy of H2 [32] matches the intersection of the extrapolated isotherm of H2 [27] and a trajectory Vtr/L + N9 (of 13.0 to 13.3 Å3/at = 90 rB3 by using the intersection of the trajectory N9 in Figure 3a and Figure 4) at around 0.5 TPa, which suggests that the transition from molecular to atomic hydrogen occurs at that pressure at 300 K.
Finally, it is noticed that Equation (1) could be used to assess the atomic volumes of polymorphs with extant structure analyses such as Ca-VII, Sc-III and -IV, Ga-IV, and metastable Ir-II to 8.0(5), 12.08(8), 10.3(11), 9.2(4), and 9.78(12) Å3/at at 170, 26, 35, 120, and 59 GPa, respectively (Table 1). The volume of hydrogen at the predicted transition at 490–530 GPa is estimated to 2.3(2) Å3/at. This and above values (marked with an asterisk) and the entire list of volumes Vtr and pressures Ptr are given in Table 1.
Table 1. Columns from left to right: Name of polymorph, pressure of phase transition at 300 K, volume of phase transition divided by the principal quantum number L, volume of phase transition at 300 K, and reference(s).
Table 1. Columns from left to right: Name of polymorph, pressure of phase transition at 300 K, volume of phase transition divided by the principal quantum number L, volume of phase transition at 300 K, and reference(s).
PhasePtr [GPa]Vtr/L [Å3/at]Vtr3/at]Reference(s)
Li-fcc7.5(1)7.6(1)15.2(2)[33,34,35]
Li-hR139.7(1.0)4.6(1)9.2(2)[34,36]
Li-c1648(2)4.48.8(4)[34,35]
Li-oC4067(2)3.4(1)6.8(2)[35]
Li-oC2497(3)3.3(2)6.6(4)[35]
Na-fcc67(1)4.10(5)12.3(30)[36]
Na-cI16115(2)3.26(1)9.77(3)[3]
Na-oP8119(2)3.15(1)9.45(6)[3]
Na-h4p200(8)2.7(1)8.1(3)[2]
K-fcc11.0(5)8.4(1)33.74(4)[37]
K-III22(43)5.64(5)22.57(3)[5]
K-IIIb 31(2)5.2(2)20.8(4)[38]
K-IV (oP8)54(2)3.69(1)14.78(4)[39]
Rb-II7.0(5)7.13(6)35.65(30)[37]
Rb-III14.3(2)6.95(1)34.77(5)[5]
Rb-IV16.7(1)5.56(1)27.8(15)[5,40]
Rb-V19.6(5)5.2(1)26.0(5)[41]
Rb-VI48.1(5)3.6(3)18.0(1.5)[41]
Rb-VII>703.4(1)17(1)[41]
Rb-VIII220(15)2.4(1)12(1)[41]
Cs-II2.37(1)8.58(1)51.47(6)[42]
Cs-III4.20(2)8.37(1)50.22(6)[5]
Cs-IV4.30(2)5.83(1)32.28(6)[43]
Cs-V10(1)5.1(2)30.6(1.2)[44]
Cs-VI72(2)2.6(2)15.6(1.2)[44,45]
Mg-bcc47(2)4.63(3)13.9(2)[46]
Ca-II19.5(15)8.02(2)32.08(8)[47,48]
Ca-III35(3)5.0(1)19.9(4)[49,50]
Ca-IV124(10)3.0(2)12.0(8)[50]
Ca-V/VI155(15)2.5(2)10.0(8)[50]
Ca-VII170 *2.0 *8this work
Sr-II3.5(2)8.30(8)41.5(4)[48]
Sr-III26.0(6)3.98(19)19.9(9)[5,49]
Sr-IV35(1)4.1(2)20.5(10)[5,49,51]
Sr-V46(3)3.5(1)17.5(5)[5,49,51]
Ba-II5.5(1)7.14(4)42.84(24)[48]
Ba-IV12.6(1)7.0(1)42.0(6)[52]
Ba-V45(2)3.5(1)21.0(6)[53]
Sc-II20(1)4.5(1)18.0(4)[54]
Sc-III104(5)3.02(2) *12.08(8)[54]
Sc-IV140(10)2.58(27) *10.3(1.1)[54]
Sc-V240(10)2.09(7)3.36(28)[19]
ω-Zr12.5(10)4.15(5)20.75(25)[55]
Zr-bcc35(2)3.5(10)17.5(50)[55,56]
Zr-bccII58–603.08(5)15.4(3)[55]
ω-Ti0(5)4.375(15)17.5(6)[20]
γ-Ti110(10)2.55(5)10.2(2)[20,21]
δ-Ti150(10)2.41(5)9.64(20)[20,21]
Ti-bcc243(10)2.07(3)8.28(12)[21]
Vrhombohedral250(20)2.02(5)8.08(20)[57]
Fe-hcp17.6(1)2.57(1)10.28(4)[58]
Os-hcpII140(10)1.88(10)11.28(60)[22]
Os-hcpIII440(20)1.5(1)9.0(6)[22]
Sn-bct10.8(1)4.60(1)23.00(5)[59]
Sn-bco32.5(10)3.93(1)19.65(5)[59]
Sn-bcc40.8(15)3.78(1)18.9(5)[59]
Pb-hcp14(1)6.95(10)41.7(6)[60]
Pb-bcc142(10)2.68(2)16.08(12)[5,61]
Ga-II2.0(2)4.52(8)18.08(32)[5]
Ga-III2.8(1)4.24(5)16.96(24)[62]
Ga-V10.5(2)4.358(5)17.432(24)[5]
Ga-IV120(10)2.3(1)*9.2(4)[62,63]
Tl-fcc3.5(5)4.3(2)25.8(1.2)[64]
Ge-II10.64.28(8)17.12(32)[65]
Ge-hp75(3)3.18(10)17.72(40)[18,66]
Ge-Cmca140(120–160) *2.68(7)10.72(28)[66]
Ge-hcp>1702.4(1)9.6(4)[18]
Diamond2.0(1)2.837(1)5.674(2)[67]
Si-II12(1)4.3(1)12.9(3)[68]
Si-V16(2)4.43(2)13.29(6)[69]
Si-VI38(2)3.83(1)11.49(3)[69]
Si-VII42(2)3.71(2)11.13(6)[69]
N, cg-phase110(7)2.58(3)5.16(6)[28]
P-A74.5(1)5.03(5)15.09(15)[70]
P-sc10(1)4.54(2)13.62(6)[70]
P-ph140(8)2.8(1)8.4(3)[70]
P-bcc280(20)2.4(1)7.2(3)[70]
As-sc25(1)4.1(1)16.4(4)[18]
As-III46(3)3.5(1)14.0(4)[18]
As-IV125(10)2.84(2)11.36(8)[18]
Sb-II8.6(2)4.42(20)22.1(1)[71]
Sb-V28(3)4.05(6)20.25(30)[71]
Bi-II2.55(1)5.30(2)31.80(12)[8]
Bi-III2.70(1)4.48(9)26.88(54)[8]
Bi-V7.7(1)4.57(10)27.42(60)[8]
S-II3.0(2)7.0(1)21.0(3)[72]
S-III38(2)4.6(1)13.8(3)[72]
S-IV103(6)3.3(1)9.9(3)[72]
S-V150(10)2.8(1)8.4(3)[72]
Se-VII14(1)4.5(1)18.0(4)[72]
Se-III26(1)4.25(7)17.0(3)[72]
Se-IV35(1)3.8(1)15.2(4)[72]
Se-V77(2)4.0(1)16.0(4)[72]
Se-VI136(7)2.87(7)11.48(28)[18]
Te-II4.0(1)5.56(1)27.80(5)[5,73]
Te-III7.4(1)5.40(1)27.00(5)[5,73]
Te-V29.2(7)4.25(2)21.25(10)[5,73]
Te-VI102(5)3.3(2)16.5(1)[74]
I-bct43(3)3.9(1)19.6(1)[75,76]
I-fcc55(4)3.78(2)18.9(1)[75,76]
Th-bct100(30)2.51(1)17.57(7)[77]
γ-Ce0.5(1)4.70(1)28.20(6)[78]
PrtI2180(10) 23.90(14)[79]
Sm-dhcp0.4(1)5.51(1)33.06(6)[80]
Sm-hR913(1)4.2(1)25.2(6)[81]
Sm-oF890(10)2.45(8)14.70(48)[81]
Nd-oF8100(10)2.30(4)13.80(24)[82]
Eu_IV31.5(20)3.32(2)19.9(1)[83]
Gd-dhcp9(3)4.3(1)26[84]
Gd-fcc26(2)3.7(1)22.2(6)[85]
Gd-dfcc33(2)3.43(10)20.5(6)[85]
Gd-VIII60.5(3)2.86(10)17.15(61)[85]
Dy-hR92.5(2)4.79(1)28.74(6)[86]
Cl-IV266(10)2.77(6)8.31(18)[29]
Ir-II591.63(2) *9.78(12)[5,87]
* Parameters calculated using Equation (1).

5. Conclusions

The volume and pressure of phase transformations of non-molecular elemental solids obeyed a universal relation of discrete corresponding states. These states, as defined in Equation (1), were close to integer multiples of the cube of the Bohr radius. Therefore, the volumes and pressures of polymorphic transitions of these solids were controlled by the number of valence electrons as specified in Equation (4). This finding was in accordance with the commonly used effective single-electron-based computational approaches for modeling elemental solids within the range of 0 to 0.5 TPa. However, Equation (1) can be used for constraining elemental metal volumes for first-principal- and empirical-potential-based calculations, thus removing the weakest point in these approaches (the proper assessment of volumes). The volumes of polymorphs at the phase transitions exhibited a general linear dependence on the pressure of the transition (see Equation (1)). Though not an equation of state, this relation establishes a very simple principle that rules over a vast range of simple and complex solid structures and a range of pressure of 0.5 TPa.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/cryst12121698/s1, Figure S1: Volume of elemental polymorphs in Å3/at divided by the principal quantum number L as function of pressure in GPa and mapped onto the set N4; Figure S2: Correlation between the normalized transformation volumes from Equation (1) and p-state occupancy.

Funding

This research received no external funding.

Data Availability Statement

All data are provided in the paper.

Acknowledgments

Comments and suggestions by S. Huang, H.-k. Mao, A. Zerr, and by two anonymous reviewers are gratefully acknowledged.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Mao, H.-K.; Chen, X.J.; Ding, Y.; Li, B.; Wang, L. Solids, liquids, and gases under high pressure. Rev. Mod. Phys. 2018, 90, 015007. [Google Scholar] [CrossRef] [Green Version]
  2. Ma, Y.M.; Eremets, M.; Oganov, A.R.; Xie, Y.; Trojan, I.; Medvedev, S.; Lyakhov, S.; Valle, M.; Prakapenka, V. Transparent dense sodium. Nature 2009, 458, 182–185. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Gregoryanz, E.; Lundegaard, L.F.; McMahon, M.I.; Guillaume, C.; Nelmes, R.J.; Mezouar, M. Structural diversity of sodium. Science 2009, 320, 1054–1057. [Google Scholar] [CrossRef]
  4. Marques, M.; McMahon, M.I.; Gregoryanz, E.; Hanfland, M.; Guillaume, C.L.; Pickard, C.J.; Ackland, G.J.; Nelmes, R.J. Crystal Structures of Dense Lithium: A Metal-Semiconductor-Metal Transition. Phys. Rev. Lett. 2011, 106, 095502. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. McMahon, M.I.; Nelmes, R.J. High-pressure structures and phase transformations in elemental metals. Chem. Soc. Rev. 2006, 35, 943–963. [Google Scholar] [CrossRef]
  6. Schwarz, U. High-pressure crystallography and synthesis. Z. Krist. 2004, 219, 943–963. [Google Scholar]
  7. Cammi, R.; Rahm, M.; Hoffmann, R.; Ashcroft, N.W. Varying Electronic Configurations in Compressed Atoms: From the Role of the Spatial Extension of Atomic Orbitals to the Change of Electronic Configuration as an Isobaric Transformation. J. Chem. Theory Comput. 2020, 16, 5047–5056. [Google Scholar] [CrossRef]
  8. Degtyareva, O.; McMahon, M.I.; Nelmes, R.J. High-pressure structural studies of group 1-5 elements. High Press. Res. 2004, 24, 319–356. [Google Scholar] [CrossRef]
  9. Winzenick, M.; Vijayakumar, V.; Holzapfel, W.B. High-pressure X-ray-diffraction on potassium and rubidium up to 50 GPa. Phys. Rev. B 1994, 50, 12381–12385. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Fabbris, G.; Lim, J.; Veiga, L.S.I.; Haskel, D.; Schilling, J.S. Electronic and structural ground state of heavy alkali metals at high pressure. Phys. Rev. B 2015, 91, 085111. [Google Scholar] [CrossRef] [Green Version]
  11. Duthie, J.C.; Pettifor, D.G. Correlation between d-band occupancy and crystal-structure in rare-earths. Phys. Rev. Lett. 1977, 38, 564–567. [Google Scholar] [CrossRef]
  12. Holzapfel, W.B. Structural systematics of 4f and 5f elements under pressure. J. Alloys Comp. 1995, 224, 319–356. [Google Scholar] [CrossRef]
  13. Eriksson, O.; Brooks, M.S.S.; Johansson, B. Orbital polarization in narrow-band systems—Application to volume collapses in light lanthanides. Phys. Rev. B 1990, 41, 7311–7314. [Google Scholar] [CrossRef]
  14. Neaton, J.B.; Ashcroft, N.W. On the constitution of sodium at higher densities. Phys. Rev. Lett. 2001, 86, 2830–2833. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Bradley, J.A.; Moore, K.T.; Lipp, M.J.; Mattern, B.A.; Pacold, J.I.; Seidler, G.T.; Chow, P.; Rod, E.; Xiao, Y.M.; Evans, W.J. 4f electron delocalization and volume collapse in praseodymium metal. Phys. Rev. B 2021, 85, 100102. [Google Scholar] [CrossRef]
  16. Iota, V.; Klepeis, J.H.P.; Yoo, C.S.; Lang, J.; Haskel, D.; Srajer, G. Electronic structure and magnetism in compressed transition metals. Appl. Phys. Lett. 2007, 90, 042505. [Google Scholar] [CrossRef]
  17. Dewaele, A.; Stutzmann, V.; Bouchet, J.; Bottin, F.; Occelli, F.; Mezouar, M. The α → ω phase transformation in zirconium followed with ms-scale time-resolved X-ray absorption spectroscopy. High Pres. Res. 2016, 36, 237–249. [Google Scholar]
  18. Akahama, Y.; Kamiue, K.; Okawa, N.; Kawaguchi, S.; Hirao, N.; Ohishi, Y. Volume compression of period 4 elements: Zn, Ge, As, and Se above 200GPa: Ordering of atomic volume by atomic number. J. Appl. Phys. 2021, 129, 02590. [Google Scholar] [CrossRef]
  19. Akahama, Y.; Fujihisa, H.; Kawamura, H. New helical chain structure for scandium at 240 GPa. Phys. Rev. Lett. 2005, 94, 195503. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Dewaele, A.; Stutzmann, V.; Bouchet, J.; Bottin, F.; Occelli, F.; Mezouar, M. High pressure-temperature phase diagram and equation of state of titanium. Phys. Rev. B 2015, 91, 134108. [Google Scholar] [CrossRef]
  21. Akahama, Y.; Kawaguchi, S.; Hirao, N.; Ohishi, Y. Observation of high-pressure bcc phase of titanium at 243GPa. J. Appl. Phys. 2020, 128, 035901. [Google Scholar] [CrossRef]
  22. Dubrovinsky, L.; Dubrovinskaia, N.; Bykova, E.; Bykov, M.; Prakapenka, V.; Prescher, C.; Glazyrin, K.; Liermann, H.-P.; Hanfland, M.; Ekholm, M.; et al. The most incompressible metal osmium at static pressures above 750 gigapascals. Nature 2015, 525, 226–228. [Google Scholar] [CrossRef] [PubMed]
  23. Tschauner, O. Observations about the pressure-dependence of ionic radii. Geoscience 2022, 12, 246. [Google Scholar] [CrossRef]
  24. Anzellini, S.; Dewaele, A.; Occelli, F.; Loubeyre, P.; Mezouar, M. Equation of state of rhenium and application for ultra high pressure calibration. J. Appl. Phys. 2014, 115, 043511. [Google Scholar] [CrossRef]
  25. Fratanduono, D.E.; Millot, M.; Braun, D.G.; Ali, S.J.; Fernandez-Pañella, A.; Seagle, C.T.; Davis, J.-P.; Brown, J.L.; Akahama, Y.; Kraus, R.G.; et al. Establishing gold and platinum standards to 1 Terapascal using shockless compression. Science 2021, 372, 1063–1068. [Google Scholar] [CrossRef]
  26. Weck, G.; Desgreniers, S.; Loubeyre, P.; Mezouar, M. Single-Crystal Structural Characterization of the Metallic Phase of Oxygen. Phys. Rev. Lett. 2009, 102, 255503. [Google Scholar] [CrossRef]
  27. Cheng, J.; Li, B.; Liu, W.; Smith, J.S.; Majumdar, A.; Luo, W.; Ahuja, R.; Shu, J.; Wang, J.; Sinogeikin, S.; et al. Ultrahigh-pressure isostructural electronic transitions in hydrogen. Nature 2019, 573, 558–562. [Google Scholar]
  28. Eremets, M.I.; Gavriliuk, A.G.; Trojan, I.A.; Dzivenko, D.A.; Boehler, R. Single-bonded cubic form of nitrogen. Nat. Mater. 2004, 3, 558–563. [Google Scholar] [CrossRef] [PubMed]
  29. Dalladay-Simpson, P.; Binns, J.; Pena-Alvarez, M.; Donnelly, M.E.; Greenberg, E.; Prakapenka, V.; Chen, X.J.; Gregoryanz, E.; Howie, R.T. Band gap closure, incommensurability and molecular dissociation of dense chlorine. Nat. Commun. 2019, 10, 1134. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Fujihisa, H.; Fujii, Y.; Takemura, K.; Shimomura, O. Structural aspects of dense solid halogens under high-pressure studied by X-ray diffraction—Molecular dissociation and metallization. J. Phys. Chem. Solids 1995, 56, 1439–1444. [Google Scholar] [CrossRef]
  31. Akahama, Y.; Miyakawa, M.; Taniguchi, T.; Sano-Furukawa, A.; Machida, S.; Hattori, T. Structure refinement of black phosphorus under high pressure. J. Chem. Phys. 2020, 153, 104704. [Google Scholar] [CrossRef] [PubMed]
  32. Li, B.; Ding, Y.; Kim, D.Y.; Wang, L.; Weng, T.-C.; Yang, W.; Yu, Z.; Ji, C.; Wang, J.; Shu, J.; et al. Probing the Electronic Band Gap of Solid Hydrogen by Inelastic X-ray Scattering up to 90 GPa. Phys. Rev. Lett. 2021, 126, 036402. [Google Scholar] [CrossRef]
  33. Hanfland, M.; Loa, I.; Syassen, K.; Schwarz, U.; Takemura, K. Equation of state of lithium to 21 GPa. Solid State Commun. 1999, 112, 123–127. [Google Scholar] [CrossRef]
  34. Hanfland, M.; Syassen, K.; Christensen, N.E.; Novikov, D.L. New high-pressure phases of lithium. Nature 2000, 408, 175–178. [Google Scholar] [CrossRef] [PubMed]
  35. Guillaume, C.L.; Gregoryanz, E.; Degtyareva, O.; McMahon, M.I.; Hanfland, M.; Evans, S.; Guthrie, M.; Sinogeikin, S.V.; Mao, H.-K. Cold melting and solid structures of dense lithium. Nat. Phys. 2011, 7, 211–214. [Google Scholar] [CrossRef] [Green Version]
  36. Hanfland, M.; Loa, I.; Syassen, K. Sodium und pressure: Bcc to fcc structural transition and pressure-volume relation to 100 GPa. Phys. Rev. B 2002, 65, 184109. [Google Scholar] [CrossRef]
  37. Olijnyk, H.; Holzapfel, W.B. Phase transitions in K and Rb under pressure. Phys. Lett. 1983, 99A, 381–383. [Google Scholar] [CrossRef]
  38. Lundegaard, L.F.; Stinton, G.W.; Zelazny, M.; Guillaume, C.L.; Proctor, J.E.; Loa, I.; Gregoryanz, E.; Nelmes, R.J.; McMahon, M.I. Observation of a reentrant phase transition in incommensurate potassium. Phys. Rev. B 2013, 88, 054106. [Google Scholar] [CrossRef]
  39. Lundegaard, L.F.; Marques, M.; Stinton, G.; Ackland, G.J.; Nelmes, R.J.; McMahon, M.I. Observation of the oP8 crystal structure in potassium at high pressure. Phys. Rev. B 2009, 80, 020101. [Google Scholar] [CrossRef]
  40. Schwarz, U.; Grzechnik, A.; Syassen, K.; Loa, I.; Hanfland, M. Rubidium-IV: A high pressure phase with complex crystal structure. Phys. Rev. Lett. 1999, 83, 4085–4088. [Google Scholar] [CrossRef]
  41. Storm, C.V.; McHardy, J.D.; Finnegan, S.E.; Pace, E.J.; Stevenson, M.G.; Duff, M.J.; MacLeod, S.G.; McMahon, M.I. Behavior of rubidium at over eightfold static compression. Phys. Rev. B 2021, 103, 224103. [Google Scholar] [CrossRef]
  42. Hall, H.T.; Merrill, L.; Barnett, J.D. A high pressure phase of cesium. Science 1964, 146, 1297–1299. [Google Scholar] [CrossRef] [PubMed]
  43. Takemura, K.; Minomura, S.; Shimomura, O. X-ray Diffraction Study of Electronic Transitions in Cesium under High Pressure. Phys. Rev. Lett. 1982, 49, 1772–1775. [Google Scholar] [CrossRef]
  44. Schwarz, U.; Takemura, K.; Hanfland, M.; Syassen, K. Crystal Structure of Cesium-V. Phys. Rev. Lett. 1998, 81, 2711–2713. [Google Scholar] [CrossRef]
  45. Takemura, K.; Shimomura, O.; Fujihisa, H. Cs(VI): A new high-pressure polymorph of cesium above 72 GPa. Phys. Rev. Lett. 1991, 66, 2014–2017. [Google Scholar] [CrossRef] [PubMed]
  46. Stinton, G.W.; MacLeod, S.G.; Cynn, H.; Errandonea, D.; Evans, W.J.; Proctor, J.E.; Meng, Y.; McMahon, M.I. Equation of state and high-pressure/high-temperature phase diagram of magnesium. Phys. Rev. B 2014, 90, 134105. [Google Scholar] [CrossRef] [Green Version]
  47. Anzellini, A.; Errandonea, D.; MacLeod, S.G.; Botella, P.; Daisenberger, D.; De’Ath, J.M.; Gonzalez-Platas, J.; Ibáñez, J.; McMahon, M.I.; Munro, K.A.; et al. Phase diagram of calcium at high pressure and high temperature. Phys. Rev. Mater. 2022, 2, 083608. [Google Scholar] [CrossRef]
  48. Anderson, M.S.; Swenson, C.A.; Peterson, D.T. Experimental equations of state for calcium, strontium, and barium metals to 20 kbar and from 4 to 295 K. Phys. Rev. B 1990, 41, 3329–3338. [Google Scholar] [CrossRef] [PubMed]
  49. Olijnyk, H.; Holzapfel, W.B. Phase transitions in alkaline earth metals under pressure. Phys. Lett. A 1984, 100, 191–194. [Google Scholar] [CrossRef]
  50. Fujihisa, H.; Nakamoto, Y.; Sakata, M.; Shimizu, K.; Matsuoka, T.; Ohishi, Y.; Yamawaki, H.; Takeya, S.; Gotoh, Y. Ca-VII: A Chain Ordered Host-Guest Structure of Calcium above 210 GPa. Phys. Rev. Lett. 2013, 110, 235501. [Google Scholar] [CrossRef]
  51. Bovornratanaraks, T.; Allan, D.R.; Belmonte, S.A.; McMahon, M.I.; Nelmes, R.J. Complex monoclinic superstructure in Sr-IV. Phys. Rev. B 2006, 73, 144112. [Google Scholar] [CrossRef] [Green Version]
  52. Nelmes, R.J.; Allan, D.R.; McMahon, M.I.; Belmonte, S.A. Self-Hosting Incommensurate Structure of Barium IV. Phys. Rev. Lett. 1999, 83, 4081–4084. [Google Scholar] [CrossRef]
  53. Kenichi, T. High-pressure structural study of barium to 90 GPa. Phys. Rev. B 1994, 50, 16238–16246. [Google Scholar] [CrossRef] [PubMed]
  54. Fujihisa, H.; Akahama, Y.; Kawamura, H.; Gotoh, Y.; Yamawaki, H.; Sakashita, M.; Takeya, S.; Honda, K. Incommensurate composite crystal structure of scandium-II. Phys. Rev. B 2005, 72, 132103. [Google Scholar] [CrossRef]
  55. Stavrou, E.; Yang, L.H.; Söderlind, P.; Aberg, D.; Radousky, H.B.; Armstrong, M.R.; Belof, J.L.; Kunz, M.; Greenberg, E.; Prakapenka, V.B.; et al. Anharmonicity-induced first-order isostructural phase transition of zirconium under pressure. Phys. Rev. B 2018, 98, 220101. [Google Scholar] [CrossRef] [Green Version]
  56. Anzellini, S.; Bottin, F.; Bouchet, J.; Dewaele, A. Phase transitions and equation of state of zirconium under high pressure. Phys. Rev. B 2020, 102, 184105. [Google Scholar] [CrossRef]
  57. Akahama, Y.; Kawaguchi, S.; Hirao, N.; Ohishi, Y. High-pressure stability of bcc-vanadium and phase transition to a rhombohedral structure at 200 GPa. J. Appl. Phys. 2021, 129, 135902. [Google Scholar] [CrossRef]
  58. Dewaele, A.; Torrent, M.; Loubeyre, P.; Mezouar, M. Compression curves of transition metals in the Mbar range: Experiments and projector augmented-wave calculations. Phys. Rev. B 2008, 78, 104102. [Google Scholar] [CrossRef]
  59. Salamat, A.; Briggs, R.; Bouvier, P.; Petitgirard, S.; Dewaele, A.; Cutler, M.E.; Cora, F.; Daisenberger, D.; Garbarino, G.; McMillan, P.F. High-pressure structural transformations of Sn up to 138 GPa: Angle-dispersive synchrotron X-ray diffraction study. Phys. Rev. B 2013, 88, 104104. [Google Scholar] [CrossRef]
  60. Kuznetsov, A.; Dmitriev, V.; Dubrovinsky, L.; Prakapenka, V.B.; Weber, H.-P. FCC–HCP phase boundary in lead. Solid State Commun. 2018, 122, 125–127. [Google Scholar] [CrossRef]
  61. Mao, H.K.; Wu, Y.; Shu, J.F.; Hu, J.Z.; Hemley, R.J.; Cox, D.E. High pressure phase transition and equation of state of lead to 238 GPa. Solid State Commun. 1990, 74, 1027–1029. [Google Scholar] [CrossRef]
  62. Degtyareva, O.; McMahon, M.I.; Allan, D.R.; Nelmes, R.J. Structural Complexity in Gallium under High Pressure: Relation to Alkali Elements. Phys. Rev. Lett. 2004, 93, 205502. [Google Scholar] [CrossRef] [PubMed]
  63. Kenichi, T.; Kobayashi, K.; Arai, M. High-pressure bct-fcc phase transition in Ga. Phys. Rev. B 1998, 58, 2482–2486. [Google Scholar] [CrossRef]
  64. Cazorla, C.; MacLeod, S.G.; Errandonea, D.; Munro, K.A.; McMahon, M.I.; Popescu, C. Thallium under extreme compression. J. Phys. Condens. Matter 2016, 28, 445401. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Menoni, C.S.; Hu, J.Z.; Spain, I.L. Germanium at high pressures. Phys. Rev. B 1986, 34, 362–368. [Google Scholar] [CrossRef] [PubMed]
  66. Takemura, K.; Schwarz, U.; Syassen, K.; Hanfland, M.; Christensen, N.E.; Novikov, D.L.; Loa, I. High-pressure Cmca and hcp phases of germanium. Phys. Rev. B 2000, 62, R10603–R10606. [Google Scholar] [CrossRef]
  67. Kennedy, C.S.; Kennedy, G.C. Equilibrium boundary between graphite and diamond. J. Geophys. Res. 1976, 81, 2467–2470. [Google Scholar] [CrossRef]
  68. McMahon, M.I.; Nelmes, R.J. New high-pressure phase of Si. Phys. Rev. B 1993, 47, 8337–8340. [Google Scholar] [CrossRef]
  69. Hanfland, M.; Schwarz, U.; Syassen, K.; Takemura, K. Crystal Structure of the High-Pressure Phase Silicon VI. Phys. Rev. Lett. 1999, 82, 1197–1199. [Google Scholar] [CrossRef]
  70. Akahama, Y.; Kawamura, H.; Carlson, S.; Le Bihan, T.; Hausermann, D. Structural stability and equation of state of simple-hexagonal phosphorus to 280 GPa: Phase transition at 262 GPa. Phys. Rev. B 2000, 61, 3139–3142. [Google Scholar] [CrossRef]
  71. Degtyareva, O.; McMahon, M.I.; Nelmes, R.J. Pressure-induced incommensurate-to-incommensurate phase transition in antimony. Phys. Rev. B 2004, 70, 184119. [Google Scholar] [CrossRef]
  72. Degtyareva, O.; Gregoryanz, E.; Mao, H.K.; Hemley, R.J. Crystal structure of sulfur and selenium at pressures up to 160 GPa. High Press. Res. 2005, 25, 17–33. [Google Scholar]
  73. Hejny, C.; Falconi, S.; Lundegaard, L.F.; McMahon, M.I. Phase transitions in tellurium at high pressure and temperature. Phys. Rev. B 2006, 74, 174119. [Google Scholar] [CrossRef]
  74. Akahama, Y.; Okawa, N.; Sugimoto, T.; Fujihisa, H.; Hirao, N.; Ohishi, Y. Coexistence of a metastable double hcp phase in bcc–fcc structure transition of Te under high pressure. Jpn. J. Appl. Phys. 2018, 57, 02560. [Google Scholar] [CrossRef]
  75. Reichlin, R.; McMahan, A.K.; Ross, M.; Martin, S. Optical, X-ray, and band-structure studies of iodine at pressures of several megabars. Phys. Rev. B 1994, 49, 3725–3733. [Google Scholar] [CrossRef] [PubMed]
  76. Fujihisa, H.; Takemura, K.; Onoda, M.; Gotoh, Y. Two intermediate incommensurate phases in the molecular dissociation process of solid iodine under high pressure. Phys. Rev. Res. 2021, 3, 033174. [Google Scholar] [CrossRef]
  77. Vohra, Y.K.; Akella, J. 5f bonding in thorium metal at extreme compressions—Phase transitions to 300 GPa. Phys. Rev. B 1991, 67, 3563–3566. [Google Scholar]
  78. Decremps, F.; Belhadi, L.; Farber, D.L.; Moore, K.T.; Occelli, F.; Gauthier, M.; Polian, A.; Antonangeli, D.; Aracne-Ruddle, C.M.; Amadon, B. Diffusionless α-γ Phase Transition in Polycrystalline and Single-Crystal Cerium. Phys. Rev. Lett. 2011, 106, 065701. [Google Scholar] [CrossRef] [Green Version]
  79. O’Bannon, E.F.; Pardo, O.S.; Soderlind, P.; Sneed, D.; Lipp, M.J.; Park, C.; Jenei, Z. Systematic structural study in praseodymium compressed in a neon pressure medium up to 185 GPa. Phys. Rev. B 2022, 105, 144107. [Google Scholar] [CrossRef]
  80. Jayarahman, A.; Sherwood, R.C. Phase transformation in samarium induced by high pressure and its effect on the antiferromagnetic ordering. Phys. Rev. 1964, 134, A691–A692. [Google Scholar] [CrossRef]
  81. Finnegan, S.E.; Pace, E.J.; Storm, C.V.; McMahon, M.I.; MacLeod, S.G.; Liermann, H.P.; Glazyrin, K. High-pressure structural systematics in samarium up to 222 GPa. Phys. Rev. B 2020, 101, 174109. [Google Scholar] [CrossRef]
  82. Finnegan, S.E.; Storm, C.V.; Pace, E.J.; McMahon, M.I.; MacLeod, S.G.; Plekhanov, E.; Bonini, N.; Weber, C. High-Pressure Structural Systematics in Neodymium to 302 GPa. Phys. Rev. B 2021, 103, 134117. [Google Scholar] [CrossRef]
  83. Husband, R.J.; Loa, I.; Munro, K.A.; McBride, E.E.; Evans, S.R.; Liermann, H.-P.; McMahon, M.I. Phase transitions in europium at high pressures. High Press. Res. 2013, 33, 158–164. [Google Scholar] [CrossRef]
  84. Golosova, N.O.; Kozlenko, D.P.; Lukin, E.V.; Kichanov, S.E.; Savenko, B.N. High pressure effects on the crystal and magnetic structure of 160Gd metal. J. Magn. Magnet. Mater. 2021, 540, 16848515. [Google Scholar] [CrossRef]
  85. Errandonea, D.; Boehler, R.; Schwager, B.; Mezouar, M. Structural studies of gadolinium at high pressure and temperature. Phys. Rev. B 2007, 75, 014103. [Google Scholar] [CrossRef]
  86. Tschauner, O.; Grubor-Urosevic, O.; Dera, P.; Mulcahy, S. Anomalous elastic behaviour in hcp-Sm-type Dysprosium. J. Phys. Chem. C 2012, 116, 2090–2095. [Google Scholar] [CrossRef]
  87. Cerenius, Y.; Dubrovinsky, L. Compressibility measurements on iridium. J. Alloys Comp. 2000, 306, 26–29. [Google Scholar] [CrossRef]
Figure 1. Volume of elemental polymorphs in Å3/at at 300 K divided by the principal quantum number L as a function of pressure in GPa. The volumes are those that are observed at the pressure of their formation through direct phase transformation from lower pressure polymorphs; the plotted pressures are the pressures of the transformation. The linear relation between these volumes and pressures and the existence of five distinct sets of pressure–volume relations is clearly visible.
Figure 1. Volume of elemental polymorphs in Å3/at at 300 K divided by the principal quantum number L as a function of pressure in GPa. The volumes are those that are observed at the pressure of their formation through direct phase transformation from lower pressure polymorphs; the plotted pressures are the pressures of the transformation. The linear relation between these volumes and pressures and the existence of five distinct sets of pressure–volume relations is clearly visible.
Crystals 12 01698 g001
Figure 2. This figure shows the data from Figure 1 with the volume corrected for the linear pressure dependence from Equation (1). The remaining constant term of each set is approximately equal to integer multiples of 2, 3, 4, 5, and 6 of 10 times the cube of the Bohr radius.
Figure 2. This figure shows the data from Figure 1 with the volume corrected for the linear pressure dependence from Equation (1). The remaining constant term of each set is approximately equal to integer multiples of 2, 3, 4, 5, and 6 of 10 times the cube of the Bohr radius.
Crystals 12 01698 g002
Figure 3. (a) Correlation between the normalized transformation volumes Vtr from Equation (1) and d-state occupancy. The volumes Ni from Equation (1) were used to define the Vtr/L sets of the alkaline metals [10] and of a selection of other elements as a function of d-state occupancy. Filled symbols: measured d-state occupancies; hollow symbols: estimated occupancies based on Equations (1) and (2). Solid lines: fitted correlation for KII, Rb-II and -IV, and Cs-II and -III. The same slope has then been applied to all data. Dashed lines give the values Ni and are guides for the eye. (b) Comparison of the atomic volumes of elemental non-molecular solids at ambient pressure and temperature with the calculated volumes Ni · L. The correlation is linear with an adjusted R2 = 0.9928 and a slope of 1.015(13), which reflects a general slight contraction of the elemental volumes relative to Ni · L.
Figure 3. (a) Correlation between the normalized transformation volumes Vtr from Equation (1) and d-state occupancy. The volumes Ni from Equation (1) were used to define the Vtr/L sets of the alkaline metals [10] and of a selection of other elements as a function of d-state occupancy. Filled symbols: measured d-state occupancies; hollow symbols: estimated occupancies based on Equations (1) and (2). Solid lines: fitted correlation for KII, Rb-II and -IV, and Cs-II and -III. The same slope has then been applied to all data. Dashed lines give the values Ni and are guides for the eye. (b) Comparison of the atomic volumes of elemental non-molecular solids at ambient pressure and temperature with the calculated volumes Ni · L. The correlation is linear with an adjusted R2 = 0.9928 and a slope of 1.015(13), which reflects a general slight contraction of the elemental volumes relative to Ni · L.
Crystals 12 01698 g003
Figure 4. Data from Figure 1 with the equations of state for gold [24], chlorine [29], hydrogen [27], and oxygen [26] shown as yellow, green, red, and blue lines, respectively. Dashed lines indicate the distinct sets that defined Vtr/L versus P as defined by Equation (1). The additional data points for hydrogen [32] and the second elastic anomaly of osmium [22] have been added. Hachured areas mark the molecular gap and the forbidden zone below Vtr/L = 10 · rB3.
Figure 4. Data from Figure 1 with the equations of state for gold [24], chlorine [29], hydrogen [27], and oxygen [26] shown as yellow, green, red, and blue lines, respectively. Dashed lines indicate the distinct sets that defined Vtr/L versus P as defined by Equation (1). The additional data points for hydrogen [32] and the second elastic anomaly of osmium [22] have been added. Hachured areas mark the molecular gap and the forbidden zone below Vtr/L = 10 · rB3.
Crystals 12 01698 g004
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Tschauner, O. Corresponding States for Volumes of Elemental Solids at Their Pressures of Polymorphic Transformations. Crystals 2022, 12, 1698. https://doi.org/10.3390/cryst12121698

AMA Style

Tschauner O. Corresponding States for Volumes of Elemental Solids at Their Pressures of Polymorphic Transformations. Crystals. 2022; 12(12):1698. https://doi.org/10.3390/cryst12121698

Chicago/Turabian Style

Tschauner, Oliver. 2022. "Corresponding States for Volumes of Elemental Solids at Their Pressures of Polymorphic Transformations" Crystals 12, no. 12: 1698. https://doi.org/10.3390/cryst12121698

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop