• Chinese Physics B
  • Vol. 29, Issue 8, (2020)
Chang Q Sun
Author Affiliations
  • School of EEE, Nanyang Technological University, Singapore639798
  • show less
    DOI: 10.1088/1674-1056/ab8dad Cite this Article
    Chang Q Sun. Rules essential for water molecular undercoordination[J]. Chinese Physics B, 2020, 29(8): Copy Citation Text show less
    O:H–O segmental length cooperative relaxation.[63,78] (a) O—O distance changes by shortening one segment and lengthening the other because of coupling interaction. The O:H always relaxes more than the H–O does because of the O:H–O segmental disparity. The primary rule for (b) O:H–O length cooperativity by compression (p), thermal excitation (t), molecular undercoordination (z), and electrostatic polarization (e).
    Fig. 1. O:H–O segmental length cooperative relaxation.[63,78] (a) O—O distance changes by shortening one segment and lengthening the other because of coupling interaction. The O:H always relaxes more than the H–O does because of the O:H–O segmental disparity. The primary rule for (b) O:H–O length cooperativity by compression (p), thermal excitation (t), molecular undercoordination (z), and electrostatic polarization (e).
    Specific-heat disparity and multiphase thermal mass density oscillation.[79] (a) Intersection points define the quasisolid phase (QS) whose boundaries close to temperatures for homogeneous ice nucleation (TN) and melting (Tm). Molecular undercoordination disperses the QS boundary outwardly by H–O contraction and O:H elongation through Einstein’s relation. (b) Segmental specific-heat ratio defines the thermal slope of density over all phases for water ice and the critical temperatures vary with volume size at the nanometer scale (T ≥ 273 K bulk water; T ≤ 273 K 1.4 nm sized droplet).[21]
    Fig. 2. Specific-heat disparity and multiphase thermal mass density oscillation.[79] (a) Intersection points define the quasisolid phase (QS) whose boundaries close to temperatures for homogeneous ice nucleation (TN) and melting (Tm). Molecular undercoordination disperses the QS boundary outwardly by H–O contraction and O:H elongation through Einstein’s relation. (b) Segmental specific-heat ratio defines the thermal slope of density over all phases for water ice and the critical temperatures vary with volume size at the nanometer scale (T ≥ 273 K bulk water; T ≤ 273 K 1.4 nm sized droplet).[21]
    Undercoordination resolved bond relaxation and O:H–O potentials.[1,82,85] (a) BOLS formulation of atomic undercoordination-resolved bond contraction (z m > 0 being the bond nature index. (b) O:H–O potential paths for the sized (H2O)N = 2 – 6 clusters (ΔdH EH > 0; ΔdL > 0, ΔEL > 0). The blue dots in (b) are the initial equilibrium for N = 6.
    Fig. 3. Undercoordination resolved bond relaxation and O:H–O potentials.[1,82,85] (a) BOLS formulation of atomic undercoordination-resolved bond contraction (z < 12), with m > 0 being the bond nature index. (b) O:H–O potential paths for the sized (H2O)N = 2 – 6 clusters (ΔdH < 0, ΔEH > 0; ΔdL > 0, ΔEL > 0). The blue dots in (b) are the initial equilibrium for N = 6.
    O:H–O bond segmental length cooperative relaxation. O—O repulsion dislocates O ions in the same direction by different amounts (insets) under (a) mechanical compression, cooling of (b) liquid and (c) quasisolid (QS) phase (in units of °C), and (d) undercoordination by reducing the (H2O)N size from N = 6 to 2. Arrows denote the master pieces and their relaxation directions. The H–O bond always relaxes less than the O:H and both of them relax contrastingly in their curvatures and slopes, irrespective of the applied stimulus or the structural order because of the O–O Coulomb repulsive coupling (reprinted with permission from Ref. [63]).
    Fig. 4. O:H–O bond segmental length cooperative relaxation. O—O repulsion dislocates O ions in the same direction by different amounts (insets) under (a) mechanical compression, cooling of (b) liquid and (c) quasisolid (QS) phase (in units of °C), and (d) undercoordination by reducing the (H2O)N size from N = 6 to 2. Arrows denote the master pieces and their relaxation directions. The H–O bond always relaxes less than the O:H and both of them relax contrastingly in their curvatures and slopes, irrespective of the applied stimulus or the structural order because of the O–O Coulomb repulsive coupling (reprinted with permission from Ref. [63]).
    Skin O:H–O segmental length and stiffness cooperative relaxation.[53] Length and stiffness (inset) transition for the (a) H–O bond and (b) O:H nonbond from the bulk value (b) to the skin (s) and to the H–O free radicals (r). Inset (a) shows the complex unit cell denoted with bulk, skin, and a vacuum slab. The P components arise from the screening and splitting of the crystal potentials by polarization.
    Fig. 5. Skin O:H–O segmental length and stiffness cooperative relaxation.[53] Length and stiffness (inset) transition for the (a) H–O bond and (b) O:H nonbond from the bulk value (b) to the skin (s) and to the H–O free radicals (r). Inset (a) shows the complex unit cell denoted with bulk, skin, and a vacuum slab. The P components arise from the screening and splitting of the crystal potentials by polarization.
    Computational H–O stretching vibration modes in the (H2O)n clusters.[88] The black dashed lines convolute the H–O vibration modes of the entire clusters. The sharp feature D corresponds to the H–O dangling bonds, C to the H–O of the O:H–O bonds between molecules at rims, features A and B to the H–O bonds inside the clusters. Reprinted with copyright permission from Ref. [88].
    Fig. 6. Computational H–O stretching vibration modes in the (H2O)n clusters.[88] The black dashed lines convolute the H–O vibration modes of the entire clusters. The sharp feature D corresponds to the H–O dangling bonds, C to the H–O of the O:H–O bonds between molecules at rims, features A and B to the H–O bonds inside the clusters. Reprinted with copyright permission from Ref. [88].
    H–O stiffness transition from the bulk to the skins of large volumes and nanodroplets.[53,74] (a) Skins of 298 K water and (253–258) K ice[65] share an identical ωH of 3450 cm−1 and the (b) D–O phonon[33] transits from below 2550 to its above for the skin of droplet. Insets show ice slipperiness, water skin toughness, and the elasticity and hydrophobicity of skins of droplet and water.
    Fig. 7. H–O stiffness transition from the bulk to the skins of large volumes and nanodroplets.[53,74] (a) Skins of 298 K water and (253–258) K ice[65] share an identical ωH of 3450 cm−1 and the (b) D–O phonon[33] transits from below 2550 to its above for the skin of droplet. Insets show ice slipperiness, water skin toughness, and the elasticity and hydrophobicity of skins of droplet and water.
    Site and orientation resolved SFG H–O vibration frequency. Undercoordination resolves the frequencies of the OB1:H–OB2 and the OB1 – H:OB2 bonds (inset) between outermost two sublayers of the ice Ih(0001) skin.[35] Insets illustrate segmental bond lengths, orientations, and frequencies of the H–O stretching vibrations. The positive peak (−1) corresponds to the H–OB2 vibration (shaded in green) and the valley (> 3270 cm−1) to the OB1-H (shaded in blue). The less coordinated OB1-H is shorter and stiffer and its H:OB2 is longer and softer than the OB1:H–OB2.
    Fig. 8. Site and orientation resolved SFG H–O vibration frequency. Undercoordination resolves the frequencies of the OB1:H–OB2 and the OB1 – H:OB2 bonds (inset) between outermost two sublayers of the ice Ih(0001) skin.[35] Insets illustrate segmental bond lengths, orientations, and frequencies of the H–O stretching vibrations. The positive peak (< 3270 cm−1) corresponds to the H–OB2 vibration (shaded in green) and the valley (> 3270 cm−1) to the OB1-H (shaded in blue). The less coordinated OB1-H is shorter and stiffer and its H:OB2 is longer and softer than the OB1:H–OB2.
    Undercoordination induced quantum entrapment and polarization.[31,63,64,67,68] (a) The O 1s energy shifts from the bulk value of 536.6 eV to 538.1 eV for the skin and to 549.7 eV for gaseous state. (b) The hydrated nonbonding electron shifts its energy from the core value of 2.4 eV and the skin of 1.2 eV to the limit of 0.4 eV for N = 5 cluster. Inset (b) shows the cluster size dependent ωH blueshift.[4]
    Fig. 9. Undercoordination induced quantum entrapment and polarization.[31,63,64,67,68] (a) The O 1s energy shifts from the bulk value of 536.6 eV to 538.1 eV for the skin and to 549.7 eV for gaseous state. (b) The hydrated nonbonding electron shifts its energy from the core value of 2.4 eV and the skin of 1.2 eV to the limit of 0.4 eV for N = 5 cluster. Inset (b) shows the cluster size dependent ωH blueshift.[4]
    Ultrafast phonon and hydrated electron processes of water droplets.[32,33] Lifetime of (a) the D–O phonons in the droplet skin compared with that in the NaBr/H2O solutions[33] and the lifetime of (b) electron hydrated by the sized (H2O)n and (D2O)n clusters and their skins.[32] Insets in (a) illustrate the ionic hydration and the core-shelled water droplet. The skin supersolidity hinders the motion dynamics of both phonons and nonbonding electrons.[55]
    Fig. 10. Ultrafast phonon and hydrated electron processes of water droplets.[32,33] Lifetime of (a) the D–O phonons in the droplet skin compared with that in the NaBr/H2O solutions[33] and the lifetime of (b) electron hydrated by the sized (H2O)n and (D2O)n clusters and their skins.[32] Insets in (a) illustrate the ionic hydration and the core-shelled water droplet. The skin supersolidity hinders the motion dynamics of both phonons and nonbonding electrons.[55]
    Supersolid skin thermal stability.[57] Temperature dependence of (a) the full-frequency Raman spectra and (b) the frequency shift of the bulk, skin, and H–O dangling bond components. The H–O dangling bond undergoes thermal expansion (redshift) at T > 300 K. Both the bulk and the skin components undergo thermal contraction, at different slopes. At T > 340 K, the skin component turns to be thermal elongation, showing the O—O repulsive weakening.
    Fig. 11. Supersolid skin thermal stability.[57] Temperature dependence of (a) the full-frequency Raman spectra and (b) the frequency shift of the bulk, skin, and H–O dangling bond components. The H–O dangling bond undergoes thermal expansion (redshift) at T > 300 K. Both the bulk and the skin components undergo thermal contraction, at different slopes. At T > 340 K, the skin component turns to be thermal elongation, showing the O—O repulsive weakening.
    Skin thermal-diffusivity elevation and specific-heat depression.[102,55] Numerical reproduction of the measured (insets) initial-temperature dependence of (a) the θ (θi, t) decay and (b) the skin-bulk temperature difference Δθ (θi, t) of warm water. High skin thermal diffusivity due to density loss ensures the characteristic intersection. (b) The Δθ (θi, t) arises from the skin lower specific heat because heat flux conserves at the interface.
    Fig. 12. Skin thermal-diffusivity elevation and specific-heat depression.[102,55] Numerical reproduction of the measured (insets) initial-temperature dependence of (a) the θ (θi, t) decay and (b) the skin-bulk temperature difference Δθ (θi, t) of warm water. High skin thermal diffusivity due to density loss ensures the characteristic intersection. (b) The Δθ (θi, t) arises from the skin lower specific heat because heat flux conserves at the interface.
    Nano-supersolid TN depression – supercooling.[48,53] Skin supersolidity takes the full responsibility for (a) water skin toughness, ice slipperiness, and (b) droplet-size and surface-curvature (inset) resolved TN depression (called supercooling).[106,112] Droplet of slightly more-curved surface freezes 68.4 s later at 269 K than the contrast deposited on smooth Ag surface.[111]
    Fig. 13. Nano-supersolid TN depression – supercooling.[48,53] Skin supersolidity takes the full responsibility for (a) water skin toughness, ice slipperiness, and (b) droplet-size and surface-curvature (inset) resolved TN depression (called supercooling).[106,112] Droplet of slightly more-curved surface freezes 68.4 s later at 269 K than the contrast deposited on smooth Ag surface.[111]
    Consequence of molecular undercoordinationBonding and electronic origin
    Thermo-mechanical anomalies1Water skin toughness.[46] The surface stress is 72.75 mJ/m2 compared with 26.6 mJ/m2 for CCl4 solution at 293 K. Surface stress drops linearly with the rise of temperature.[47]Supersolidity of higher melting point Tm and lower freezing temperature TN and polarization.
    2Slipperiness of ice.[9,48] The slipperiness of wet surfaces is most for hydrophilic/hydrophobic contact but least for hydrophilic/hydrophilic interaction.[49] Ice on ice has a higher friction coefficient.O:H nonbond high elastic adaptivity and surface dipolar repulsivity.
    3Hydrophobicity and elasticity.[50,51] Water droplet dances rounds before merging into the bulk when falling on to liquid water.Skin elastic repulsive supersolidity.
    4Ice skin premelting and nanorheology.[6,52] The skin is viscoelastic over a wide span of temperature. with a viscosity up to two orders of magnitude larger than pristine water.Gel-like, viscoelastic supersolid phase.
    5Skin low mass density.[53] The skin mass density is confirmed 0.75 g/cm3 opposing to classical thermodynamics prediction, XRD revealed 5.9% skin O—O elongation with respect to 2.8 Å length or 15.6% density loss at 298 °C. In contrast, the skin O—O for liquid methanol contracts by 4.6% associated with a 15% density gain.[54]O:H expands more than H–O contraction.
    6Thermal diffusivity and specific heat.[55] High thermal diffusivity and low specific heat ensure heat outward flow and high surface temperature in the thermal transport of warm water to cold drain – Mpemba effect.[56]Skin density-loss dominance of diffusivity and Debye temperature offset by phonon frequency relaxation.
    7Thermal stability.[17,57] Raman H–O phonon skin-component at 3450 cm−1 is less sensitive to temperature than its bulk at 3200 cm−1. Skin and bulk components undergo thermal contraction yet the dangling H–O bond undergoes thermal expansion.Heating can hardly deform further the undercoordination deformed H–O bond.
    8Nanobubble durability and reactivity.[58–61] Nanobubble is mechanically and thermally endurable and chemically more reactive.Skin supersolidity ensures mechanical and thermal stability; polarization raises the reactivity.
    9Supercooling and superheating.[51,62] Water nanodroplet or bubbles undergo superheating at melting and supercooling at freezing and evaporating, whose extent is droplet size dependence. A 1.2 nm sized droplet freezes at temperature below 172 K and the monolayer skin melts at 320 K.QS boundary dispersion by O:H–O relaxation through Einstein’s relation: ΔΘx ∝ Δωx, raising the Tm and lowering the TV and TV.
    Electron phonon characteristics10Electron entrapment and polarization.[31,63,64] O 1s core-level shifts from the bulk value of 536 eV to 538 eV for the skin and to 540 eV for the gaseous state. The hydrated nonbonding electron shifts its bound energy from the bulk value of 2.4 eV for the interior and 1.2 eV for the skin to the limit of 0.4 eV when the cluster size is reduced to N = 5.H–O bond contraction deepens the local potential well, which entraps the core levels; densely entrapped core electrons polarizes the nonbonding electrons.
    11Phonon stiffness.[53,65] Skins of 298 K water and (253–258) K ice share an identical H–O phonon frequency of 3450 cm−1, in contrast to the bulk values of 3200(water) and 3150(ice) cm−1 and 3650 cm−1 for the H2O monomer in gaseous phase. H–O dangling bond frequency of 3610 cm−1. H–O phonon frequency increases linearly with the inverse of cluster size.[4]Phonon frequency shift is proportional to the square-root of segmental cohesive energy and inversely to the segmental length.
    12Refractive index.[66] The refractive index of water (measured at λ = 589.2 nm) skin is higher than it is in the bulk.Polarization dominance of dielectric permittivity.
    13Lifetime of skin H–O phonon[33] and hydrating electrons.[32,64,67–69] Skin hydrated electrons and stiffened phonons have longer lifetime or slower relaxation dynamics.Skin polarization and boundary wave reflection; quasi standing wave formation.
    Length-energy transition14Bond length.[54,63] H–O bond contracts from 1.00 Å to 0.95 Å and the O:H from 1.70 Å to 1.95 Å; H–O dangling bond length of 0.9 Å. Skin dOO was measured as 2.965 Å compared to the bulk water of 2.70 Å.[54]
    15Bond energy.[63,70] The O:H–O segmental energies transit from (0.2, 4.0) to (0.1, 4.6) eV when moving from the bulk to the skin in the least-coordinated gaseous H–O bond energy of 5.10 eV. Gaseous H–O dissociation requires 121.6 nm laser beam irradiation.[70]
    Table 1. Typical examples for the molecular undercoordination resolved anomalies of low-dimensional water ice.
    Chang Q Sun. Rules essential for water molecular undercoordination[J]. Chinese Physics B, 2020, 29(8):
    Download Citation