Water's phase diagram: From the notion of thermodynamics to hydrogen-bond cooperativity

https://doi.org/10.1016/j.progsolidstchem.2015.03.001Get rights and content

Abstract

This presentation features recent progress in understanding the phase diagram of water and ice from the perspective of hydrogen bond (O:H–O) cooperative relaxation with focus on how the segmental length and the containing angle of the O:H–O bond change with mechanical compression and thermal excitation. By interplaying theoretical predictions, numerical computations, and phonon spectrometrics, we firstly examined the relaxation dynamics of O:H–O bond segmental length and phonon stiffness of: i) liquid water at 300 K and ice at 80 K as a function of pressure, ii) liquid water cooling from 350 K to 80 K under the ambient pressure, iii) mechanical freezing of the ambient water under compression up to 1.83 GPa, and, iv) liquid water heating from 253 to 753 K under 30 MPa pressure. Observations allow us to classify the TC(P) phase boundaries of water and ice into four types according to their slopes and then formulate them in terms of hydrogen bond relaxation in segmental length and containing angle. Observations reinforce the essentiality and effectiveness of hydrogen bond notion in dictating the unusual behavior of water and ice and clarify the bonding dynamics during phase transition, which is beyond the scope of classical thermodynamics.

Graphical abstract

Hydrogen bond (O:H–O) relaxation in segmental length and its containing angle dictates respectively phase boundaries in water's phase diagram.

  1. Download : Download high-res image (215KB)
  2. Download : Download full-size image

Introduction

Water is ubiquitously important to all lives and it is too strange, too anomalous, and too challenge. Understanding the anomalous behavior of water and ice remains incomprehensive despite extensive dedications in past decades. Alternative ways of thinking and approaching are indispensable though the classical thermodynamics and the contemporary quantum mechanics have contributed significantly. Consideration is necessary from the perspective of the asymmetrical, local, and short-range inter- and intra-molecular interactions within the hydrogen bond (O:H–O) and the involvement of Coulomb repulsion between electron pairs on oxygen ions [1].

Appendix shows the phase diagram of water and ice. Ice formed by water behaves strangely at lower temperatures and higher pressures. Water–H2O – seems to be a simple molecule: two hydrogen atoms connected to a central oxygen atom in a V-shape. In everyday ice cooled from liquid water, which scientists call Ice Ih, the water molecules line up in a hexagonal pattern; this is why snowflakes all have six-sided patterns (The “h” stands for hexagonal). A variation called Ice Ic at lower temperatures, found in ice crystals floating high up in the atmosphere, forms cubic crystals, in fogs and clouds. The crystal structure of the ice is fairly loose – the reason that ice Ih is less dense than liquid water – and the O:H intermolecular nonbonds become longer and weaker but the intramolecular H–O bonds become shorter and stiffer than they are in liquid water.

At higher pressures, the usual hexagonal structure breaks down, and the bonds rearrange themselves in more compact, denser crystal structures, neatly labeled with Roman numerals: Ice II, Ice III, Ice IV and so on. Scientists have also discovered several forms of ice in which the water molecules are arranged randomly, as in glass. At a pressure of about 200 MPa, Ice Ih turns into a different type of crystalline ice, Ice II. Ice II does not occur naturally on Earth. Even at the bottom of the thickest portions of the Antarctic ice cap, the weight of three miles of water and ice pushes down at only one-quarter of the pressure necessary to make Ice II. But planetary scientists expect that Ice II, and possibly some other variations, like Ice VI (around 1 GPa ∼104 atm pressure), exist inside icier bodies in the outer solar system, like the Jupiter moons Ganymede and Callisto.

With pressure high enough, the temperature need not even be cold for ice to form. Scientists considered what happens to tectonic plates after they are pushed back down into Earth's interior. At about 100 miles down, the temperature of these descending plates is 300–400 °C – well above the boiling point of water at the surface – but cool compared with that of surrounding rocks. The pressure of 2 GPa at this depth could be great enough to transform any water that was there into a solid phase known as Ice VII. No one knows whether ice can be found inside Earth, because no one has yet figured out a way to look 100 miles underground. Just as salt melts ice at the surface, other molecules mixing with the water could impede the freezing.

Ice also changes its form with dropping temperatures. In hexagonal ice, the usual form, the oxygen atoms are fixed in position, but the O:H nonbonds between water molecules are continually breaking and reattaching, tens of thousands of times a second. At temperatures cold enough – below −200 °C – the nonbonds freeze as well, and normal ice starts changing into Ice XI (orthorhombic structure). Astronomers were probably already looking at Ice XI on the surface of Pluto and on the moons of Neptune and Uranus. But instruments currently are not sensitive enough to distinguish the slight differences among the ices.

From ice XII to ice XVI, found just a decade ago are also with many new features. For instance [2], the cadge-structured ice XVI has a density of 0.81 g/cm3 as the stable low-temperature phase of water at negative pressures (that is, under tension). This hollow hydrate structure exhibits cooling expansion below about 55 K, and that it is mechanically more stable and has at low temperatures larger lattice constants than the filled hydrate. When pressure is increased to 60 GPa, ice X phase forms with identical H–O and O:H length of 1.1 Å, corresponding to density of 1.84 times that of water at 4 °C [3], [4].

Ice can turn to be partially ionic (2H2O → H3Oδ+(one lone pair) + HOδ−(δ = 0.62; three lone pairs), which is realized by exchanging H with an electron at extremely high pressure (2 TPa) and high temperature (2000 K) [5]. The occurrence of this ionic phase follows the break-up of the typical O–H covalently bonded tetrahedrons in the hydrogen symmetric atomic phases and is originated from the volume reduction favorable for a denser structure packing.

The contribution to the lattice energy from the O:H intermolecular interaction increases and that from the intramolecular H–O bonding decreases when pressure is increased up to 2 GPa, as calculated using first principle and Quantum Mote Caro calculations [6]. The elegantly used TIPnP (n varies from 1 to 5) model series and the TIP4Q/2005 modeled water ice but they exclude the possibility of bond relaxation and charge polarization. These rigid non-polarizable models can hardly reproduce the anomalies of water ice with high satisfaction [7], [8].

One often connects the structure and property Q of a substance directly to the external stimuli such as pressure, volume, entropy, and temperature Q(PV, ST,…), which is exactly that the classical thermodynamics deals with large trunks. Such treatment employs concepts of entropy, enthalpy, Gibbs and Helmholtz free energy, and so on so forth with standard deviation δ depending on the sample size N in the form of σ ∝ N−1/2. This approach provides little information on how the chemical bond responds to external stimulus.

A number of formulas exists describing the TC(P) phase boundaries of water ice from the classical thermodynamic point of view, for the Liquid–Vapor phase transition. Unfortunately, few approaches are available for formulating the boundaries in general cases. For instances, Clausius–Clapeyron equation [9] describes water vapor under typical atmospheric conditions (near standard temperature and pressure) and August–Roche–Magnus formula [10] approximates the temperature dependence of the saturation vapor pressure PS:(dPsdT=Lv(T)PsRvT2(ClausiusClapeyron)Ps(T)=6.1094exp(17.625TT+243.04)(AugustRocheMagnus)where Lv is the specific latent heat of evaporation of water and Rv is the gas constant of vapor.

However, Pauling [11] indicated that the nature of the chemical bond bridges the structure and the property of a substance, and therefore, one can modulates the Q by relaxing the chemical bonds through controlling external stimuli such as pressure, temperature, coordination environment, chemical composition, electric and magnetic field, etc [12]. The latter approach considers only the change of the order, length, and energy of interatomic bond under the quoted excitations. Pauling's premise challenges us to revisit the physical nature of a complex system by decomposing it into a simple representative of all bonds involved. Therefore, understanding how the water molecule works at transition from the perspective of O:H–O bond relaxation in segmental length, containing angle is important in understanding how water interacts with all the biological molecules in living organisms.

This presentation aims to accounting recent progress in due respect with focus on revisiting the TC–P phase diagram of water and ice from the perspective of O:H–O bond cooperative relaxation under mechanical compression and thermal excitation. By interplaying with theoretical predictions, numerical computations, and phonon spectrometrics, we firstly examined the O:H–O bond segmental length and phonon stiffness of: i) liquid water at 300 K and ice at 80 K as a function of pressure, ii) liquid water cooling from 350 K to 80 K under the ambient pressure, iii) mechanical freezing of the ambient water under compression up to 1.83 GPa, and, iv) liquid water heating from 253 to 753 K under 30 MPa pressure. Observations allow us to classify the TC(P) phase boundaries of water and ice into four types according to their slopes and then to formulate them in terms of hydrogen bond relaxation in segmental length and containing angle.

This exploration revealed the following: i) H–O bond elongation dictates the negatively-sloped TC(P) boundaries for the VII–VIII and the Liquid–Quasisolid phase transition; ii) O:H nonbond contraction dominates the positively-sloped TC(P) profiles for the Liquid–Vapor phase transition; and, iii) O:H–O containing angle relaxation governs those of zero-sloped (TC = constant at the Ic – XI boundary) or infinitely-sloped (PC = constant at the (XII, XIII)-X boundary) TC(P) profiles. Numerical reproduction of the negatively-sloped TC(P) curves results in the H–O cohesive energy 3.97 eV for water and ice and duplication of the TC(P) boundary for Liquid–Vapor transition (called Regelation discovered by Faraday in 1859) turns out the pressure trend of the O:H length change. Raman examination revealed that the O:H and H–O contract simultaneously when turning the ambient water into ice VI by compression.

Section snippets

Dimer bond: an isolated oscillator

Generally, one can focus on the representative bond that is an average of all in a substance from the perspective of Fourier transformation that correlated the distribution in real space and behavior in the energy or momentum domains. The length and energy of this representative bond determines the structural and performance of the substance [12].

Fig. 1a shows a pairing potential u(r) for the dimer bond in a regular substance, being taken as an isolated oscillator in the first order

Mechanical compression of ice

Molecular dynamics calculations [3] transformed the measured V(P) profile [25] of ice into the dx(P) curves (x = H for H–O bond shorter than 1.1 Å and x = L for O:H nonbond longer than 1.1 Å) shown in Fig. 2a. The extrapolated dx(P) curves meet at the point of proton centralization occurring in phase X at O–O distance of 2.20 Å under 59 GPa compression [26], [27], which clarifies that the dH = dL in phase X [26] arises from the dL contraction and dH elongation instead of proton quantum

TC(P) classification and formulation

The phase diagram of water and ice in Appendix (Fig. 8) shows the following boundary TC(P) features according to their slopes:TC(P)=(δ(PC)Constf1(P)f2(P))dTC(P)dP=((XXI,X(VII,VIII),etc)0(IcVI,XVVI,etc)>0(LiquidVapour,Liquid(III,IV,V,VII))<0(VIIVIII,IhLiquid,etc)

δ(PC) is a Kronig function. If P = PC, δ(PC) = 1, else δ(PC) = 0, the critical PC keeps constant irrespective of temperature. Const indicates that TC does not change with P. The f1,2(P) is a certain type of P dependent

Outlook

Phase diagram of water and ice provides direct information of O:H–O bond relaxation along the boundaries of phase transition. Mechanical compression shortens and stiffens the O:H nonbond and lengthens and softens the H–O bond in most phases towards proton centralization. Thermal excitation changes the segmental length irregularly according to their relative specific heat. Reproduction of the negatively-sloped TC(P) profiles along the Liquid–Quasisolid and the VII-VIII boundaries confirmed that

Funding

Financial support received from NSF (Nos.: 21273191, 91227202) China is acknowledged.

Notes

The authors declare no competing financial interest.

References (49)

  • Y. Wang et al.

    High pressure partially ionic phase of water ice

    Nat Commun

    (2011)
  • B. Santra et al.

    Hydrogen bonds and van der Waals forces in ice at ambient and high pressures

    Phys Rev Lett

    (2011)
  • C. Vega et al.

    Simulating water with rigid non-polarizable models: a general perspective

    PCCP

    (2011)
  • J. Alejandre et al.

    A non-polarizable model of water that yields the dielectric constant and the density anomalies of the liquid: TIP4Q

    PCCP

    (2011)
  • K. Wark

    Generalized thermodynamic relationships in the thermodynamics

    (1988)
  • O. Alduchov et al.
  • L. Pauling

    The nature of the chemical bond

    (1960)
  • C.Q. Sun
    (2014)
  • X.J. Liu et al.

    Coordination-resolved spectrometrics of atomistic local bonding and electronic dynamics

    Chem Rev

    (2015)
  • G.R. Desiraju

    A bond by any other name

    Ang Chem Int Ed

    (2011)
  • L.R. Falvello

    The hydrogen bond, front and center

    Ang Chem Int Ed

    (2010)
  • Y. Huang et al.

    Hydrogen-bond asymmetric local potentials in compressed ice

    J Phys Chem B

    (2013)
  • Y. Huang et al.

    Size, separation, structure order, and mass density of molecules packing in water and ice

    Sci Rep

    (2013)
  • X. Zhang et al.

    A common supersolid skin covering both water and ice

    PCCP

    (2014)
  • Cited by (76)

    • Hydrogen-bonded structures and low temperature transitions of the confined water in subnano channels

      2023, Spectrochimica Acta - Part A: Molecular and Biomolecular Spectroscopy
    • Studies on micromorphology and permeability of hydrate-bearing porous media

      2023, Fuel
      Citation Excerpt :

      The diagram was divided into three main areas: liquid, solid, and vapor areas. Notably, in the solid area, orthorhombic ice, cubic ice (Ic), and hexagonal ice (Ih) were formed where the temperature was below 0 °C and the pressure was under 212.9 MPa [22] in a pure water system. In this study, Ic, Ih, and sphere ice were observed.

    • Perturbative vibration of the coupled hydrogen-bond (O:H–O) in water

      2022, Advances in Colloid and Interface Science
      Citation Excerpt :

      The ωH blueshift and the accompanied ωL redshift disperses the specific heat curves, lowering the TN and TV while raising the Tm, known as superheating at melting and supercooling at the freezing and evaporating points, as well as the known phenomenon of surface premelting [28]. The strong polarization entitled the skins high elasticity and hydrophobicity, made ice slippery and water skin tough [79]. A water droplet bounces repeatedly or persists on water surface [51].

    View all citing articles on Scopus
    View full text