You are on page 1of 10

Article

Liquid–liquid transition and critical point in


sulfur

https://doi.org/10.1038/s41586-020-2593-1 Laura Henry1, Mohamed Mezouar1 ✉, Gaston Garbarino1, David Sifré1, Gunnar Weck2 &
Frédéric Datchi3
Received: 24 April 2019

Accepted: 26 May 2020


The liquid–liquid transition (LLT), in which a single-component liquid transforms into
Published online: 19 August 2020
another one via a first-order phase transition, is an intriguing phenomenon that has
Check for updates
changed our perception of the liquid state. LLTs have been predicted from computer
simulations of water1,2, silicon3, carbon dioxide4, carbon5, hydrogen6 and nitrogen7.
Experimental evidence has been found mostly in supercooled (that is, metastable)
liquids such as Y2O3–Al2O3 mixtures8, water9 and other molecular liquids10–12. However,
the LLT in supercooled liquids often occurs simultaneously with crystallization,
making it difficult to separate the two phenomena13. A liquid–liquid critical point
(LLCP), similar to the gas–liquid critical point, has been predicted at the end of the LLT
line that separates the low- and high-density liquids in some cases, but has not yet
been experimentally observed for any materials. This putative LLCP has been invoked
to explain the thermodynamic anomalies of water1. Here we report combined in situ
density, X-ray diffraction and Raman scattering measurements that provide direct
evidence for a first-order LLT and an LLCP in sulfur. The transformation manifests
itself as a sharp density jump between the low- and high-density liquids and by distinct
features in the pair distribution function. We observe a non-monotonic variation of
the density jump with increasing temperature: it first increases and then decreases
when moving away from the critical point. This behaviour is linked to the competing
effects of density and entropy in driving the transition. The existence of a first-order
LLT and a critical point in sulfur could provide insight into the anomalous behaviour
of important liquids such as water.

The pressure–temperature (P–T) phase diagram of sulfur exhibits discontinuous change of density associated with this process. So far,
important similarities to that of phosphorus, which is so far the only no in situ structural or vibrational studies have been conducted in the
element for which a direct in situ realization of an LLT has been unam- pressure region below 3 GPa in the mixed molecular–polymeric liquid.
biguously demonstrated14–16. The stable sulfur solids at ambient pres- We performed in situ X-ray absorption, X-ray diffraction and Raman
sure (100 kPa), α- and β-sulfur17,18, consist of S8 molecules, whereas at scattering measurements at the beamline ID27 of the European Syn-
high pressure and temperature, the stable polymorph is a polymeric chrotron Radiation Facility (ESRF)26 to probe the density, structure
solid composed of helical chains19. At room pressure the molecular and dynamics evolution of liquid sulfur in the P–T domain at 0–3 GPa
character is conserved in the liquid upon melting at 388 K and up to and 300–1,100 K (see Supplementary Information sections S1–S3 for
432 K, where the so-called ‘λ-transition’ occurs20. The λ-transition has the methods). The P–T paths are presented in the experimental phase
been described as a ‘living’ polymerization transition21, reversible and diagram of sulfur in Fig. 1.
incomplete (the polymer content reaches a maximum of ~60% at the Density measurements were obtained using a Paris–Edinburgh press
boiling point, T = 718 K), in which a fraction of the S8 cyclic molecules along eight isothermal (P1–P8 in Fig. 1) and two isobaric (P9, P10) path-
open up and coalesce into long polymeric chains or rings. It is asso- ways. The accuracy of the density measured by this method is of the
ciated with a large increase in viscosity and an anomalous, but not order of 1% (Supplementary Information section S1). X-ray diffrac-
discontinuous, density variation20,22. For P > 5 GPa and T > 1,000 K, tion patterns of the sample were also collected at each P–T point to
several P–T domains with different thermal and electrical properties confirm that the sample was fully molten. As shown in Fig. 2a, below
have been proposed in liquid sulfur23. An experimental study24 reported 1,000 K, along isothermal pathways P1–P5, we systematically observed
at pressures above 6 GPa the existence of a purely polymeric liquid a discontinuous jump in density over a very narrow pressure range
composed of long chains below 1,000 K, which split to shorter chains of ~0.07 GPa, which strongly suggests the existence of a first-order
at higher temperatures. Ab initio molecular dynamics simulations25 phase transition between a low- (LDL) and a high-density liquid (HDL).
reproduced this chain breakage in the compressed liquid but found no Discontinuous density shifts were also observed upon varying the

1
European Synchrotron Radiation Facility (ESRF), Grenoble, France. 2CEA, DAM, DIF, Arpajon, France. 3Institut de Minéralogie, de Physique des Matériaux et de Cosmochimie (IMPMC), Sorbonne
Université, CNRS UMR 7590, MNHN, Paris, France. ✉e-mail: mezouar@esrf.fr

382 | Nature | Vol 584 | 20 August 2020


P8

Liquid P7
P6
1,000 Cp
P5

LDL HDL

Temperature (K)
P4
800
P3
P10
P2
II
600 E P11
III
P1
C

I
AB D
P9 Solid
400

0 0.5 1.0 1.5 2.0 2.5 3.0


Pressure (GPa)

Fig. 1 | Phase diagram of sulfur around the LLT. P1–P8: isothermal pathways A, B, C, D and E (blue filled triangles) along path P11 indicate the P, T conditions
followed during the density measurements presented in Fig. 2. P1, P2, P4–P7 of the selected X-ray diffraction data in Fig. 3. I, II and III are the (P, T) points of
were made on compression, whereas P3 (diamonds) and P7 (open black circles) the Raman spectra presented in Fig. 3. The black dashed line is the transition
were made on decompression. For clarity, P7 and P8 are shown up to 3 GPa only. line between the LDL domain (yellow) and the HDL domain (pink) that
P9 and P10 are isobaric pathways followed during the density measurements terminates at the critical point Cp (black solid circle).
presented in Supplementary Fig. 9 (Supplementary Information section S1).

temperature at constant load along pathways P9, P10 (Supplementary As shown in ref. 29, the third peak is a fingerprint of the S8 molecule
Fig. 9). These density jumps were accompanied by sudden changes in because it occurs at the average distance of the third and fourth neigh-
the structure factor S(Q) of the sulfur melt, as shown in Fig. 2b. This is bours in an S8 ring, as deduced from the structure of the molecular
particularly apparent in the width and position of the first diffraction α-sulfur crystal. When the temperature is increased in the LDL domain
peak, which change abruptly at the transition. The density variation above the λ-transition (points B at 0.17 GPa, 442 K and C at 0.36 GPa,
estimated from the measured S(Q) using the methodology of ref. 27 487 K), the observed evolution is also very similar to that described in
(see Supplementary Information section S3) compares very well with the literature for the ambient-pressure liquid28. Namely, the positions
that derived from the X-ray absorption measurements (see inset of and intensities of the first- and second-neighbour peaks are weakly
Fig. 2c), giving an independent confirmation of the discontinuous affected, whereas the third peak is strongly reduced in intensity and
density jump at the transition. becomes bimodal. These changes in the third- and fourth-neighbour
Figure 2c presents X-ray radiographic images taken along a com- distribution are a signature of the rapid increase of polymer content
pression pathway at 980 K and at pressures of 1.6–2.5 GPa. These and the associated reduction of the S8 content above the λ-transition.
images show that below (i) and above (iv) the transition, the sample Indeed, the peak at 4.45 Å, characteristic of S8 molecules, is reduced
is homogeneous, whereas at the transition an interface separating a in intensity, and a new component, originating from the formation
‘bubble’ of HDL from the surrounding LDL appears (ii and iii). As seen in of long polymeric chains or rings, appears at 4 Å and grows with tem-
Supplementary Video 1, the HDL bubble grows as the load is increased, perature. The similarities between the present PDF in the region from
until the sample is fully in the HDL phase. These observations provide 4 Å to 5 Å and those reported in ref. 28 from ambient-pressure neutron
compelling evidence of the coexistence of the LDL and HDL phases diffraction can be appreciated from the comparison between the inset
at the transition and, together with the density and structure factor of Fig. 3a and Fig. 3b. We note, however, that the new component in the
measurements, confirm the first-order nature of the LLT. ambient-pressure PDF appears at a larger distance, around 4.2 Å, which
At 1,090 K and 1,100 K (pathways P7 and P8 in Fig. 1), we did not is probably due to the lower density of the ambient-pressure liquid.
observe (within uncertainties) any discontinuous shift of the density We now come to the structural modifications in the PDF across the
as a function of pressure; this indicates the presence of an LLCP. The LDL–HDL transition. As seen in Fig. 3c and Supplementary Informa-
LLCP was probably crossed along pathway P6 at ~2.15 GPa and 1,035 K tion section S15, no change (within uncertainties) occurs on the first
(star in Fig. 2a), where the density measurements show clear anoma- and second peak positions, showing that the S–S bond length and the
lies (see Supplementary Information section S2), whereas at lower ⟨S–S–S⟩ angle are the same as in the LDL. The most important modifica-
and higher pressures along this isotherm, the density appears to tions occur again in the third- and fourth-neighbour distributions. The
vary continuously with pressure. As shown in Fig. 2d, we observed a bimodal shape of the third peak is maintained but the component at
non-monotonic evolution of the density discontinuity with tempera- 4.45 Å is even more reduced, and the component located at 4 Å in the
ture: starting from zero at the LLCP, it first increases to a maximum of LDL undergoes a sudden shift in position to 4.15 Å in the HDL. This shows
~7.5% at about 750 K, and then decreases. that the local order in the liquid changes at the transition, and further
Figure 3a shows the pair distribution function (PDF) g(r) obtained by suggests that the polymer content in the HDL is larger than in the LDL.
Fourier transform of the measured S(Q) at five selected points along The latter point is confirmed by the comparison of the Raman spectra
the P8 pathway: A, B and C are in the LDL domain, whereas D and E are measured in the LDL and HDL shown in Fig. 3d. In the HDL domain, we
in the HDL domain, close to the LDL–HDL transition line (see Fig. 1). The observe an increase in intensity of the stretching mode at 460 cm−1 that
LDL PDF at point A (0.11 GPa, 428 K) is very similar to those reported for is assigned to the polymeric chains30, concomitant with a decrease of
the ambient-pressure molecular liquid below the λ-transition28,29. It has the molecular bending mode at 152 cm−1 and the breathing mode at
three well defined peaks at 2.05(2), 3.39(2) and 4.45(2) Å (uncertainties 220 cm−1 of the S8 molecule, attesting that the latter are residual in
indicate 1 s.d.), in very good agreement with the previous cited works. the HDL region.

Nature | Vol 584 | 20 August 2020 | 383


Article
a c
1.20 550 K (P1) i
1.20 650 K (P2)
1.15 845 K (P4)
950 K (P5)
1.10 1,035 K (P6)
1,090 K (P7)
1.15 1.05 1,100 K (P8) ii
Critical point
1.00
Density variation

1.5 2.0 2.5 3.0 3.5


1.10 Pressure (GPa)

iii

1.05

iv
1.00

0.5 1.0 1.5 2.0


Pressure (GPa)

b 0.64 GPa d
0.68 GPa
0.85 GPa
1.2 0.98 GPa 8

1.0
6
Density difference (%)
Structure factor

0.8
X-ray absorption
PDF
4
Density variation

1.05
0.6

2
1.00
0.4

0.6 0.8 1.0


Pressure (GPa)
0.2 0
10 20 30 40 50 600 700 800 900 1,000
Q (nm–1) Temperature (K)

Fig. 2 | First-order LLT in sulfur. a, Relative pressure variation of the liquid data collected on decompression at 740 K are shown in Extended Data Fig. 2.
density, ρ/ρ 0 (ρ 0 is the density of the lowest pressure point for each isotherm) The variation of the density calculated from the associated PDF (black filled
collected along seven isothermal pathways (P1, P2 and P4–P8 in Fig. 1). For squares) is presented in the inset together with the one obtained from the
clarity, the density jump obtained on decompression along P3 and along the direct density measurements (red empty squares). c, X-ray radiography of
isothermal paths P9 and P10 are presented in Extended Data Figs. 1, 3. Main liquid sulfur across the LDL–HDL transition line at T = 980 K and at pressures
panel: at temperatures below 1,030 K, a clear density jump is observed along all between 1.6 GPa and 2.5 GPa. Results are shown for pure LDL (i), LDL and HDL
the isothermal paths. At ~1,035 K, a density anomaly is detected in the vicinity coexistence (ii, iii), pure HDL (iv). The yellow arrows indicate the LDL–HDL
of the LLCP (see Extended Data Fig. 4). Above the LLCP (inset), a continuous boundary. d, Temperature evolution of the density jump. The black and blue
variation of the density is observed. b, Structure factors S(Q) of liquid sulfur symbols correspond to the isothermal pathways (P1–P8) and the red symbols to
collected along the isothermal path P2 (T = 650 K). The red arrow emphasizes the isobaric (P9, P10) pathways. The maximum of 7.5% (± 1 s.d.) is located at
the shift of the first peak position of S(Q) at the LDL–HDL transition. The S(Q) ~750 K. Error bars indicate 1 s.d.

This work thus demonstrates that sulfur undergoes a first-order The entropy reduction across the LLT may be due in part to the increase
phase transition between two thermodynamically stable liquids, with in polymer content revealed by these experiments and the associated
clear experimental evidence of a sharp density increase and structural reduction in the mixing entropy. However, the observed changes in
modifications. We stress that this LLT is distinct from the long-known the PDF also indicate that the local conformation of neighbouring
λ-transition, which is associated with second-order-like changes in polymeric units is modified to a more compact arrangement imposed
density20,22 and heat capacity31. Furthermore, the λ-transition tempera- by the density increase, leading to a reduction in the conformation
ture slowly decreases with pressure32, whereas for the present LLT the entropy as well.
transition temperature increases with pressure (see also Supplemen- Because of the shape of the transition line and the presence of a
tary Information section S4). By virtue of the Clapeyron equation, and critical point, this LLT in sulfur strongly resembles the well known
because a positive jump of density occurs at the transition, this indi- liquid–gas transition. However, there is an important difference: the
cates that the entropy of the HDL is smaller than that of the LDL, which non-monotonic variation of the density jump with temperature of the
contrasts with the entropy increase in the LDL at the λ-transition33. LLT, which first increases from zero as the temperature is decreased

384 | Nature | Vol 584 | 20 August 2020


a b 1.4 d
423 K
A 473 K
4 1.3 1.3 573 K
1.2
B

PDF
1.2
C
PDF
1.1
1.2 GPa, 600 K (III)
1.1
D 1.0
3 E
1.0
0.40 0.42 0.44 0.46 0.48 0.50
0.40 0.45 0.50 r (nm)
r (nm) 4.6 4.6
c

Raman intensity
4.5 4.5
0.56 GPa 571 K (E)
PDF

2
4.4 4.4
0.36 GPa 487 K (D)
4.3 4.3

Distance (Å)
0.21 GPa 487 K (C)
Third/fourth neighbours 1 bar, 595 K (II)
4.2 4.2
0.17 GPa 442 K (B)
4.1 4.1
1
0.11 GPa 428 K (A) 4.0 4.0

3.9 3.9
3.5 Second neighbours 3.5
1 bar, 394 K (I)
3.4 3.4
0
3.3 3.3
0.1 0.2 0.3 0.4 0.5 0.6 0.1 0.2 0.3 0.4 0.5 0.6 100 200 300 400 500 600
r (nm) Pressure (GPa) Raman shift (cm–1)

Fig. 3 | Local order in the LDL and HDL sulfur. a, The PDF, g(r), of liquid data of ref. 28 at 423 K, 473 K and 573 K. The curves were reproduced with
sulfur at selected pressure and temperature conditions along path P11 of permission from ref. 28 (American Physical Society, 1990). c, Positions of the
Fig. 1. Curves A–E correspond to the P, T conditions in Fig. 1 at which the second, third and fourth neighbours as a function of pressure, as deduced from
measurements were performed. The curves are vertically offset by 0.3 for the peak positions of the PDF along path P11. d, Selected Raman spectra of
clarity. The inset shows a magnification of the PDF in the region 0.4–0.5 nm, liquid sulfur collected at the (P, T) points I, II and III of Fig. 1: I, LDL below the
which contains contributions from third and fourth neighbours. λ-transition; II, LDL above the λ-transition; III, HDL. Error bars indicate 1 s.d.
b, Magnification of the third peak of g(r) from the ambient-pressure neutron

from the LLCP, and then decreases, in contrast to the monotonic homogeneous-nucleation temperature. Finally, an LLCP has also been
increase of density along the liquid–gas transition. Such a behaviour predicted in simple molecular systems, such as H2 (ref. 6) and N2 (ref.
7
was recently predicted by ab initio molecular dynamics calculations ), but their experimental observation remains extremely challenging.
along the LLT line of phosphorus34, which suggested that the order The LLCP in sulfur, being in a P–T range easily accessible by experiment,
parameter describing the LLT contains contributions from both the provides a unique opportunity for the study of critical phenomena
density and the entropy and that at least at low temperatures, entropy— associated with LLTs. We thus expect that the present work will generate
rather than density—governs the transition. This is at odds with the a new interest in LLTs that will provide a solid basis for understanding
liquid–gas transition, for which density is the sole order parameter. the principles that govern LLTs in general. Future studies should also
A two-order-parameter model including density and a bond-order focus on deciphering the microscopic processes at the origin of the
parameter describing locally favoured structures has been proposed13 LLT in sulfur.
to explain the existence of LLTs. The putative LLT in water has also been
described as entropy-driven, on the basis of a model in which water
is considered as an ‘athermal solution’ of two molecular structures Online content
with different entropies and densities35. We note, however, that the Any methods, additional references, Nature Research reporting sum-
present LLT in sulfur is different from that in water and phosphorus, maries, source data, extended data, supplementary information,
in the sense that the transition line has a positive slope in sulfur but acknowledgements, peer review information; details of author con-
a negative one in water and phosphorus. This may signal that sulfur tributions and competing interests; and statements of data and code
belongs to a different class of LLT. availability are available at https://doi.org/10.1038/s41586-020-2593-1.
This work also provides the first, to the best of our knowledge, experi-
mental evidence for a critical point terminating the line of an LLT. Such
an LLCP was proposed in phosphorus at about 3,500 K and 0.02 GPa 1. Poole, P. H., Sciortino, F., Essmann, U. & Stanley, H. E. Phase behaviour of metastable
(ref. 34), conditions that have not yet been achieved experimentally. water. Nature 360, 324–328 (1992).
2. Harrington, S., Zhang, R., Poole, P. H., Sciortino, F. & Stanley, H. E. Liquid–liquid phase
In supercooled liquid silicon, classical empirical calculations have transition: evidence from simulations. Phys. Rev. Lett. 78, 2409–2412 (1997).
predicted an LLCP at negative pressures (−0.6 GPa, 1,120 K)36 but ab 3. Sastry, S. & Angell, C. A. Liquid–liquid phase transition in supercooled silicon. Nat. Mater.
initio calculations recently determined that the HDL– and HDL–vapour 2, 739–743 (2003).
4. Boates, B., Teweldeberhan, A. M. & Bonev, S. A. Stability of dense liquid carbon dioxide.
spinodals form a continuous reentrant curve, making supecooled Si Proc. Natl Acad. Sci. USA 109, 14808–14812 (2012).
a critical-point-free system37. In water, the existence of an LLCP was 5. Glosli, J. N. & Ree, F. H. Liquid–liquid phase transformation in carbon. Phys. Rev. Lett. 82,
proposed1 in 1992 to explain the many anomalies in the thermody- 4659–4662 (1999).
6. Morales, M. A., Pierleoni, C., Schwegler, E. & Ceperley, D. M. Evidence for a first-order
namic properties of water, such as the heat capacity, compressibility liquid–liquid transition in high-pressure hydrogen from ab initio simulations. Proc. Natl
and thermal expansion coefficients, and evidence for this LLCP has Acad. Sci. USA 107, 12799–12803 (2010).
been explored and debated ever since. The experimental observa- 7. Boates, B. & Bonev, S. First-order liquid–liquid phase transition in compressed nitrogen.
Phys. Rev. Lett. 102, 015701 (2009).
tion of this hypothetical LLCP in water may never be possible as it 8. Aasland, S. & McMillan, P. F. Density-driven liquid–liquid phase separation in the system
is located in the ‘no man’s land’, that is, the P−T domain below the Al2O3–Y2O3. Nature 369, 633–636 (1994).

Nature | Vol 584 | 20 August 2020 | 385


Article
9. Mishima, O. & Stanley, H. E. The relationship between liquid, supercooled and glassy 25. Plašienka, D., Cifra, P. & Martoňák, R. Structural transformation between long and
water. Nature 396, 329–335 (1998). short-chain form of liquid sulfur from ab initio molecular dynamics. J. Chem. Phys. 142,
10. Woutersen, S., Ensing, B., Hilbers, M., Zhao, Z. & Angell, C. A. A liquid–liquid transition in 154502–154512 (2015).
supercooled aqueous solution related to the HDA-LDA transition. Science 359, 1127–1131 26. Mezouar, M. et al. Development of a new state-of-the-art beamline optimized for
(2018). monochromatic single-crystal and powder X-ray diffraction under extreme conditions at
11. Tanaka, H., Hurita, R. & Mataki, H. PRL 92, Liquid–liquid transition in the molecular liquid the ESRF. J. Synchrotron Rad. 12, 659–664 (2005).
triphenyl phosphite. Phys. Rev. Lett. 92, 025701–025704 (2004). 27. Eggert, J., Weck, G., Loubeyre, P. & Mezouar, M. Quantitative structure factor and density
12. Kurita, R. & Tanaka, H. On the abundance and general nature of the liquid–liquid measurements of high-pressure fluids in diamond anvil cells by X-ray diffraction: argon
phase transition in molecular systems. J. Phys. Condens. Matter 17, 293–302 and water. Phys. Rev. B 65, 174105 (2002).
(2005). 28. Bellissent, R., Descotes, L., Boué, F. & Pfeuty, P. Liquid sulfur: local-order evidence of a
13. Murata, K. & Tanaka, H. Microscopic identification of the order parameter governing polymerization transition. Phys. Rev. B 41, 2135–2138 (1990).
liquid–liquid transition in a molecular liquid. Proc. Natl Acad. Sci. USA 112, 5956–5961 29. Vahvaselkä, K. S. & Mangs, J. M. X-Ray diffraction study of liquid sulfur. Phys. Scr. 38,
(2015). 737–741 (1988).
14. Katayama, Y. et al. First-order liquid–liquid phase transition in phosphorus. Nature 403, 30. Kalampounias, A. G., Kastrissios, D. T. & Yannopoulos, S. N. Structure and vibrational
170–173 (2000). modes of sulfur around the λ-transition and the glass transition. J. Non-Cryst. Solids
15. Monaco, G., Falconi, S., Crichton, W. A. & Mezouar, M. Nature of the first-order phase 326–327, 115–119 (2003).
transition in fluid phosphorus at high temperature and pressure. Phys. Rev. Lett. 90, 31. Braune, H. & Moller, O. The specific heat of liquid sulfur. Z. Naturforsch. B 9a, 210–217
255701 (2003). (1954).
16. Katayama, Y. et al. Macroscopic separation of dense fluid phase and liquid phase of 32. Kuballa, M. & Schneider, G. Differential thermal analysis under high pressure I:
phosphorus. Science 306, 848–851 (2004). investigation of the polymerisation of liquid sulfur. Ber. Bunsenges. Phys. Chem 75,
17. Steudel, R. & Eckert, B. Solid sulfur allotropes. Top. Curr. Chem. 230, 1–80 (2003). 513–516 (1971).
18. Templeton, L. K., Templeton, D. H. & Zalkin, A. Crystal structure of monoclinic sulfur. 33. Steudel, R. Liquid sulfur. Top. Curr. Chem. 230, 81–116 (2003).
Inorg. Chem. 15, 1999–2001 (1976). 34. Zhao, G. et al. Anomalous phase behavior of first-order fluid–liquid phase transition in
19. Crichton, W. A., Vaughan, G. B. M. & Mezouar, M. In situ structure solution of helical sulfur phosphorus. J. Chem. Phys. 147, 204501 (2017).
at 3 GPa and 400C. Z. Kristallogr. 216, 417–419 (2001). 35. Holten, V. & Anisimov, M. A. Entropy-driven liquid–liquid separation in supercooled water.
20. Sauer, G. E. & Borst, L. B. Lambda transition in liquid sulfur. Science 158, 1567–1569 Sci. Rep. 2, 713 (2012).
(1967). 36. Vasisht, V. V., Saw, S. & Sastry, S. Liquid–liquid critical point in supercooled silicon.
21. Tobolsky, A. V. & Eisenberg, A. Equilibrium polymerization of sulfur. J. Am. Chem. Soc. 81, Nat. Phys. 7, 549–553 (2011).
780–782 (1959). 37. Zhao, G. et al. Phase behavior of metastable liquid silicon at negative pressure: ab initio
22. Zheng, K. M. & Greer, S. C. The density of liquid sulfur near the polymerization molecular dynamics. Phys. Rev. B 93, 140203 (2016).
temperature. J. Chem. Phys. 96, 2175–2182 (1992).
23. Brazhkin, V. V., Popova, S. V. & Voloshin, R. N. Pressure–temperature phase diagram of Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in
molten elements: selenium, sulfur and iodine. Physica B 265, 64–71 (1999). published maps and institutional affiliations.
24. Liu, L. et al. Chain breakage in liquid sulfur at high pressures and high temperatures.
Phys. Rev. B 89, 174201 (2014). © The Author(s), under exclusive licence to Springer Nature Limited 2020

386 | Nature | Vol 584 | 20 August 2020


the figures produced by L.H. with contributions from all the co-authors. The manuscript was
Data availability written by M.M. and F.D. with contributions from all the co-authors.

The data that support the findings of this study are available from the Competing interests The authors declare no competing interests.
corresponding author upon request. Source data are provided with
Additional information
this paper.
Supplementary information is available for this paper at https://doi.org/10.1038/s41586-020-
2593-1.
Acknowledgements We acknowledge the European Synchrotron Radiation Facility for Correspondence and requests for materials should be addressed to M.M.
provision of synchrotron beamtime at beamline ID27, the Agence Nationale de la Recherche Peer review information Nature thanks Yoshio Kono, Wenge Yang and the other, anonymous,
for financial support under grant number ANR 13-BS04-0015 (MOFLEX) and Almax easy Lab for reviewer(s) for their contribution to the peer review of this work.
providing the diamond cylinders. Reprints and permissions information is available at http://www.nature.com/reprints.

Author contributions The original idea was conceived by M.M. The experiments were
performed by L.H., G.G., D.S. and M.M. with equal contributions. The data were analysed and
Article

Extended Data Fig. 1 | Density discontinuity at 740 K. a, Raw datasets of The black arrow indicates the density jump. b, Resulting isothermal density
isothermal X-ray absorption profiles I/I0 (where I0 and I are the incident and curve of sulfur (red) and density variation of NaCl pressure standard (blue).
transmitted intensities, respectively) collected on decompression at 740 K. Error bars indicate 1 s.d.
Extended Data Fig. 2 | Structure factors. Structure factors (S(Q)) of liquid
sulfur collected on decompression along the isothermal path at T = 740 K.
Article

Extended Data Fig. 3 | Isothermal density discontinuity. Density of liquid sulfur as a function of temperature along isobaric paths P9 at 0.4 GPa (left) and P10 at
1.3 GPa (right). Error bars indicate 1 s.d.
Extended Data Fig. 4 | LLCP in sulfur. a, b, X-ray absorption profiles I/I0 in the at temperatures below (c; 950 K) and above (d; 1,090 K) the critical point. The
horizontal (a) and vertical (b) directions in the vicinity of the critical point. red arrow in c indicates the I/I0 discontinuity at the LLT. No I/I0 discontinuity is
During the measurements, the X-ray beam was stopped by the upper and lower observed at temperatures above the critical point (d).
anvils of the Paris–Edinburgh press. c, d, Horizontal X-ray absorption profiles

You might also like