ArticlePDF Available

THz spectroscopy of liquid H2O and D2O

Authors:

Abstract and Figures

We have measured and analyzed the dielectric (0.1–2 THz) response of liquid H2O and D2O from 270 to 368 K. The response has been modeled using a Debye model with a fast and a slow decay time. By shifting the temperature scale for the slow decay time of D2O by 7.2 K we find identical behavior for D2O and H2O. The temperature dependence and isotope shift of the intermolecular structural relaxation characterized by the slow decay time can be modeled with a singular point at 228 K for H2O and 235 K for D2O.
Content may be subject to copyright.
VOLUME 82, NUMBER 14 PHYSICAL REVIEW LETTERS 5A
PRIL 1999
THz Spectroscopy of Liquid H2Oand D2O
Cecilie Rønne,1Per-Olof Åstrand,2and Søren R. Keiding1
1Department of Chemistry, Aarhus University, Langelandsgade 140, DK-8000 Århus C, Denmark
2Condensed Matter Physics and Chemistry Department, Risø National Laboratory, DK-4000 Roskilde, Denmark
(Received 15 May 1998)
We have measured and analyzed the dielectric (0.12 THz) response of liquid H2O and D2O from
270 to 368 K. The response has been modeled using a Debye model with a fast and a slow decay time.
By shifting the temperature scale for the slow decay time of D2O by 7.2 K we find identical behavior for
D2O and H2O. The temperature dependence and isotope shift of the intermolecular structural relaxation
characterized by the slow decay time can be modeled with a singular point at 228 K for H2O and 235 K
for D2O. [S0031-9007(99)08896-1]
PACS numbers: 61.25.Em, 77.22.d
The structure and dynamics of liquid water constitute
a central theme in contemporary natural science [110].
In biological systems, liquid water defines the environ-
ment for biological activity by supporting and mediating
biochemical reactions. Understanding the physical and
chemical properties of liquid water, and, in particular, the
intermolecular hydrogen bonds, is therefore often consid-
ered a prerequisite for understanding biology and chem-
istry in aqueous solutions on a molecular level. Liquid
water and ice possess many unusual properties compared
with the majority of liquids and solids, and it is often spec-
ulated whether the unique role of water in defining bio-
logical activity is caused by these anomalous properties.
Consequently, liquid water is perhaps the most studied
chemical system, with numerous experimental and theo-
retical studies [1]. In spite of these efforts, the properties
of liquid water are still far from being understood at a
molecular level. For instance, large isotope effects are
seen in some properties, such as the temperature of maxi-
mum density, which occur at 277.2 K in H2Osldand at
284.4 K in D2Osld[11], while other properties, such as
the static dielectric constant, show little difference for the
two isotopes [12,13].
The relaxational response due to intermolecular fluctua-
tions of polar liquids can be observed in the microwave
and far-infrared (FIR) region of the electromagnetic
spectrum. We have previously used THz time-domain
spectroscopy (TDS) to measure the complex dielectric
function of H2Osldin the region from 100 GHz to 2 THz
(3 to 67 cm21) as a function of temperature from a su-
percooled state (271 K) to near the boiling point (366 K)
at ambient pressure [14]. In agreement with a study
combining several microwave techniques sn,100 GHzd
with conventional FIR spectroscopy sn.176 GHzd[15],
we observed that the time correlation function of the
macroscopic polarization of liquid water is well described
by a biexponential decay, with a Debye relaxation time
stD.2psdas well as a fast st2.50 fsdrelaxation time.
A description of the TDS technique, a careful examina-
tion of the H2Oslddata and data processing together with
a comparison of the resulting relaxation times with relax-
ation times obtained in various spectroscopic studies as
well as molecular dynamics simulations can be found in
Ref. [14]. In this Letter we present the results from an
equivalent investigation of D2Osld. We focus on the tem-
perature dependence and the isotope shift of the H2Osld
and D2Osldrelaxation times and discuss the dielectric re-
laxation in view of the prevailing models of the tempera-
ture dependence of liquid water properties.
In 1976 Speedy and Angell constructed a model, where
the temperature dependence of several anomalous thermo-
dynamic properties of H2Osldcould be extrapolated from
a singular point at 228 K [16]. Many glass forming liq-
uids experience diverging structural relaxation times when
supercooled. This gives rise to a temperature dependence
for the structural relaxational predicted by the mode cou-
pling theory [7] to t~jT2T
S
j
2g, where the structural
relaxational will diverge at TS. Accordingly, the observa-
tions of Speedy and Angell have inspired a wealth of theo-
retical and experimental investigations focusing on the
behavior of supercooled/metastable water as the origin of
the anomalous properties of water at ambient temperature
and pressure. Among the water models, recently reviewed
in Ref. [10], two involve a singular/critical temperature
where the thermodynamic response functions of water di-
verge. One is the “stability limit” hypothesis that assumes
a spinodal temperature line in the P-Tphase diagram be-
yond which the liquid phase becomes mechanically unsta-
ble [17]. The other is the “liquid-liquid phase transition”
hypothesis [4,1820]. This model assumes a line of co-
existence between two liquid phases, a low-density liquid
(LDL) phase at the low-pressure side and a high-density
liquid (HDL) phase at the high-pressure side. This co-
existence terminates in a second critical point, C0, whose
location has been suggested to be at temperatures below
0±C and at either a slightly negative pressure [21] or at
a high positive pressure [1820]. However, as discussed
by Mishima and Stanley [10], the location of a critical
point in a model will be extremely dependent on the pa-
rameters used in the model as, for example, the simula-
tion potential. Furthermore, Poole et al. [18] have shown
that their two-component model can reproduce both a
2888 0031-9007y99y82(14)y2888(4)$15.00 © 1999 The American Physical Society
VOLUME 82, NUMBER 14 PHYSICAL REVIEW LETTERS 5A
PRIL 1999
spinodal line and the second critical point depending on
the parameters used. Generally, two-component models
provide a simple way of accounting for many thermo-
dynamic anomalies of liquid water. A common feature
among the most recent two-component models is that wa-
ter can be considered exhibiting a two-state temperature-
dependent equilibrium between a LDL with high local
tetrahedral ordering and a HDL [18,20,22]. An alternative
approach is based on percolation theory and contains thus
no singular point [23,24]. The main experimental diffi-
culty in establishing a consistent phase diagram is that ho-
mogeneous crystallization occurs before the hypothesized
point of divergence of the response functions is reached,
and in order to determine if there really exists a singular
point experiments very close to TSare required. Accord-
ingly, it has not yet been possible to definitively decide
which is the preferred model. In the present investigation
of the temperature dependence of the dielectric relaxation
we use the isotope shift between H2Osldand D2Osldto
examine how the two relaxation processes are related to
the liquid structure. The slow relaxation process is found
to be the relaxation of the liquid structure, and this process
is analyzed in terms of the existing models.
The experiments utilized reflection mode TDS, where
ultrashort THz pulses were reflected from a Si window
in the sample cell containing the liquid. Each THz pulse
would be divided into a pulse reflected from the front of
the silicon window and a delayed pulse reflected from the
silicon-liquid interface. Measurement of the frequency-
dependent change in phase and amplitude of the second
pulse with respect to the first pulse permitted the complex
dielectric function of the liquid to be obtained over the
entire 0.1 to 2 THz spectral range of the THz pulses [14].
The dielectric function of H2Osldat 315 K is shown in
Fig. 1 as circles. The amplitude spectrum of the THz
pulse used in the experiment is shown overlaid as crosses.
The imaginary part of the dielectric function increases by
approximately 30% when the temperature is raised from
271 to 367 K, and the real part is most sensitive to the
temperature changes at frequencies below 0.5 THz. A
Debye model including two relaxation times, a slow stDd
and a fast st2d,
ˆ´svd´`1´S
1
11ivtD
1´1
`
11ivt2
,(1)
has previously been found to give a satisfying fit of
the complex dielectric function of H2Osld[25] between
273 and 303 K [15]. The slow relaxation process has
been found to have simple Debye character down to
221 ±C [26]. We find that the dielectric response for
both H2Osldand D2Osldbetween 271 and 368 K is
successfully represented by a biexponential model (solid
line in Fig. 1). The resulting relaxation times show that
both the slow [Fig. 2(a)] and the fast [Fig. 2(b)] processes
for H2Osldand D2Osldspeed up as the temperature is
raised. The slow process is faster in H2Osldthan in
FIG. 1. The imaginary and the real part of the dielectric
constant for H2Osldat 315 K is shown ssdin (a) and (b),
respectively. The solid line is a fit to the double Debye
model described in Eq. (1), where the static dielectric constant,
´SsTf±Cgd, was constrained to 87.91e20.00458Tand 78.25f12
4.617 31023Tp11.22 31025T2
p22.7 31028T3
pgwith
TpsT225d, for H2Osld[12] and D2Osld[13], respectively,
and the individual data points were weighted according to the
amplitude spectrum of the THz pulse s3dgiving less weight to
the low and high frequency limits of the spectrum. The contri-
butions from the two modes, with relaxation strength ´S
1
for the slow process stDdand ´1
`for the fast process
st2d, are shown with dashed and dotted lines, respectively.
D2Osldat a given temperature. In contrast, within the
experimental uncertainty no isotope shift is observed for
the fast process. The temperature dependence of Debye
relaxation times obtained in this work is very similar
to what is observed in previous microwave studies [27]
[e.g., inset of Fig. 2(a)]. Since our discussion of the
experimental data will be based on the relation between
the temperature dependence of the relaxation times for the
two isotopes a small systematic deviation observed in the
numerical values will not influence the conclusions.
We have previously shown that the temperature de-
pendence of tDcan be correlated with the viscosity via
the Einstein-Stoke-Debye relation [14], thus the Debye
process corresponds to a relaxation of the liquid struc-
ture and not to a single-molecule motion. Assuming that
the temperature of maximum density (TMD) can be used
as a fingerprint for the structure, we can explore the role
of temperature dependence of the liquid structure. For
instance, it has been noted that properties of H2Osldand
D2Osldthat depend solely on the liquid structure, e.g.,
density, show a constant isotope ratio if they are shifted
2889
VOLUME 82, NUMBER 14 PHYSICAL REVIEW LETTERS 5A
PRIL 1999
FIG. 2. Temperature dependence of the slow (a) and fast (b)
relaxation time in H2Osldand D2Osld. The error bars indicate
the 95% confidence limit. The inset in (a) shows single Debye
model relaxation times, tMW , obtained by Yastremskii [27]
using microwave spectroscopy.
with the TMD [28]. Since the TMD for H2Osldand
D2Osldis separated by 7.2 K at atmospheric pressure, we
have compared tDfor H2Osldat Twith tDfor D2Osldat
T27.2 K [Fig. 3(a)]. We find that the slow relaxation
times are equal within the experimental uncertainty for the
isotopes on this “shifted” temperature scale. A qualita-
tive background for shifting the temperature according to
TMD can be found within the two-component model [28]:
The temperature-dependent mole fraction, fsTd, describ-
ing the partition between the LDL sfdand HDL s12fd
water structures, can be assumed to be equal for the two
isotopes at the density maximum, fT277.2
H2OfT284.4
D2O.
Thus, according to the different TMD the hypothesized
singular temperature, TS288 K for H2Osld, will corre-
spond to TS235.2 K for D2Osld. We have correspond-
ingly plotted lnstDdvs 2lnsTyTS21din Fig. 3(b) for
both isotopes. We observe the same linear dependence for
both H2Osldand D2Osldwith g1.57s0.03dfor a linear
fit to the merged isotope data set. This is in good agree-
ment with the temperature dependence observed using
microwave relaxation times of H2Osldwith TS228 K
determined to g1.55 (Ref. [29]) or g1.791s0.02d
(Ref. [16]). The observed power law dependence is not
sufficient to decide among the current water models, but
the observed connection between TMD and TSmakes
clear that a proper model of liquid water should provide a
combined explanation of both the isotope and temperature
FIG. 3. (a) Temperature dependence of the slow relaxation
times in H2Osldand D2Osldcompared on a shifted scale
where the D2Osldrelaxation times measured at Tare shown
at temperature T27.2 K. (b) Angell plot of the slow re-
laxation times, lntDplotted versus 2lnsTyTS21d, where
TS228 K (235.2 K) for H2OsldfD
2
Osldg. The linear fit,
lntD0.11s0.03d11.57s0.03dx, shows the power depen-
dence. The inset in (b) shows the Yastremskii [27] relax-
ation times, tMW , plotted the same way. Also tMW for H2Osld
and D2Osldcoincidence when TSis shifted by 7.2 K [inset of
Fig. 3(b)].
dependence. We find that a two-component model with
a liquid-liquid phase transition [30] is consistent with our
observations. The temperature-dependent equilibrium be-
tween HDL and LDL provides a simple rationale for the
TMD and power dependence observed for both isotopes,
with a difference in singular temperature corresponding to
the difference in TMD.
We have performed a comparison between the fast re-
laxation times of H2Osldat Tand D2Osldat T27.2 K,
but contrary to the slow relaxation times, we do not
find a correlation within the experimental uncertainty.
This indicates that the fast relaxation process is not re-
lated to fluctuations in the liquid structure, but rather
to a molecular relaxation process. Furthermore, we ob-
served that the relative contribution to the fast relax-
ation time, C2s´1
`
dys´S
`
d, increases from
approximately 2.5%(0.3%) to 4%(1%) from the lowest to
the highest temperature in the measurement. Recently,
Woutersen et al. [8] found two relaxation times for the
rotational anisotropy in a femtosecond study of the vibra-
tional spectrum of DOH in D2Osld. They interpreted the
2890
VOLUME 82, NUMBER 14 PHYSICAL REVIEW LETTERS 5A
PRIL 1999
observations in terms of a two-component (HDL/LDL)
model and assigned the fast decay time of 0.7 ps to the
LDL component. In contrast, the slow decay time of
13 ps was assigned to both states. Our results are consis-
tent with the results of Woutersen et al. [8]. The relative
contribution of the fast relaxation process is indeed in-
creasing with increasing temperature, which is consistent
with that of the relative contribution of the LDL com-
ponent increases with temperature [10]. Futhermore, the
overall small contribution from the fast relaxation process
indicates that the Debye relaxation is present in both com-
ponents. A small contribution from the fast relaxation
process to the LDL component may thus be important for
a model of the temperature dependence of liquid water.
To summarize, we have measured and analyzed the
dielectric response function for both H2Osldand D2Osld
as a function of temperature between 0.1 and 2 THz.
The dielectric relaxation was successfully represented
by a biexponential model with a fast s,300 fsdand a
slow s.2psddecay time. The temperature dependence
and isotope shift observed in this experiment for the
intermolecular relaxation characterized by slow decay
times are consistent with structural relaxation, where the
structural properties of H2Osldand D2Osldbecome equal
when compared on a temperature scale shifted by 7.2 K
for D2Osld. Furthermore, the temperature dependence of
the slow relaxation times can be modeled from a singular
point at 228 K for H2O and 235 K for D2O.
[1] Water: A Comprehensive Treatise, edited by F. Franks
(Plenum Press, New York, 19721981), Vols. 17.
[2] C.A. Angell, Annual Review of Physical Chemistry
(Annual Reviews, Palo Alto, 1983), Vol. 34, p. 593.
[3] C. A. Angell, Nature (London) 331, 206 (1988).
[4] P. H. Poole, F. Sciortino, U. Essmann, and H.E. Stanley,
Nature (London) 360, 324 (1992).
[5] N. Agmon, J. Phys. Chem. 100, 1072 (1996).
[6] H. Tanaka, Nature (London) 380, 328 (1996).
[7] Supercooled Liquids, edited by J.T. Fourkas, D. Kivelson,
U. Mohanty, and K.A. Nelson (American Chemical
Society, Washington, 1997), Vol. 676.
[8] S. Woutersen, U. Emmerichs, and H. J. Bakker, Science
278, 658 (1997).
[9] O. Mishima and H. E. Stanley, Nature (London) 392, 164
(1998).
[10] O. Mishima and H. E. Stanley, Nature (London) 396, 329
(1998).
[11] G. S. Kell, J. Chem. Eng. Data 12, 66 (1967).
[12] H. Zaghloul and H. A. Buckmaster, J. Phys. D 18, 2109
(1985).
[13] Gmelins Handbuch (Verlag Chemie, Weinham, 1964),
Lieferung 6.
[14] C. Rønne, L. Thrane, P.-O. Åstrand, A. Wallqvist, K. V.
Mikkelsen, and S.R. Keiding, J. Chem. Phys. 107, 5319
(1997).
[15] H. J. Liebe, G.A. Hufford, and T. Manabe, Int. J. Infrared
Millim. Waves 12, 659 (1991).
[16] R.J. Speedy and C. A. Angell, J. Chem. Phys. 65, 851
(1976).
[17] R. J. Speedy, J. Phys. Chem. 86, 982 (1982).
[18] P. H. Poole, F. Sciortino, T. Grande, H. E. Stanley, and
C.A. Angell, Phys. Rev. Lett. 73, 1632 (1994).
[19] F. Sciortino, P. H. Poole, U. Essmann, and H.E. Stanley,
Phys. Rev. E 55, 727 (1997).
[20] C. T. Moynhian, Mater. Res. Soc. Symp. Proc. 455, 411
(1997).
[21] H. Tanaka, in Supercooled Liquids, edited by J. T.
Fourkas, D. Kivelson, U. Mohanty, and K.A. Nelson
(American Chemical Society, Washington, DC, 1977).
[22] M.-P. Bassez, J. Lee, and G. W. Robinson, J. Phys. Chem.
91, 5818 (1987).
[23] H.E. Stanley and J. Teixeira, J. Chem. Phys. 73, 3404
(1980).
[24] S. Sastry, P. Debenedetti, F. Sciortino, and H. E. Stanley,
Phys. Rev. E 53, 6144 (1996).
[25] J. T. Kindt and C.A. Schmuttenmaer, J. Phys. Chem. 100,
10373 (1996).
[26] D. Bertolini, M. Cassettari, and G. Salvetti, J. Chem. Phys.
76, 3285 (1982).
[27] P. S. Yastremskii, J. Struct. Chem. 12, 483 (1971).
[28] M. Vedamuthu, S. Singh, and G. W. Robinson, J. Phys.
Chem. 100, 3825 (1996).
[29] C. A. Angell, in (Ref. [1]), Vol. 7.
[30] C. A. Jeffery and P. H. Austin, J. Chem. Phys. 110, 484
(1999).
2891
... The production of terahertz (THz) radiation is receiving increasing interest in a wide range of domains, such as homeland security 1-3 , medical diagnosis 4-6 , art conservation 7-10 or archeology 11,12 . THz waves can propagate through a few millimeters of optically opaque materials, provided the water concentration is low 13,14 , and can be used to identify complex molecules by coherent THz spectroscopy [15][16][17][18] . Other notable applications of THz waves include control of leaf water content [19][20][21] , plasma excitation 22 , electron acceleration 23 , high harmonic (sideband) generation [24][25][26] , or the fabrication of ferromagnetic structures 27 . ...
... cylindrical plasma of length L and radius a. Here, the only mechanism generating THz waves is the photocurrent density excited by an LP-P laser pulse, and thus, according to Eq. (14), the local THz yield evolves as sin ϕ(z). Derived from the radiated electric field in the far-field approximation, You et al. obtained the angular THz intensity spectrum (see Appendix for a detailed derivation): ...
Preprint
Full-text available
The spatial and spectral distributions of terahertz (THz) pulses emitted by two-color air plasmas are theoretically investigated for focused laser pulses and in the filamentation regime. We derive a so-called ''augmented'' conical emission model, which, similarly to the one originally proposed by You et al.\ [Phys.\ Rev.\ Lett.\ {\bf 109}, 183902 (2012)], involves phase matching between laser harmonics along the plasma channel, the plasma density and length, and the emitted frequency as key parameters. Our augmented model, however, treats envelope effects and accounts for transverse variations of both plasma- and Kerr-driven potential THz emitters. We highlight the importance of the characteristic spatio-spectral distributions of these two conversion mechanisms in the expression of the angular radiated power. The results of our model are successfully compared with data provided by a comprehensive, fully space and time-resolved unidirectional solver. Importantly, these numerical simulations clear up the effective plasma length along which THz emission develops, compared with the dephasing length along which the laser fundamental and second harmonic become out-of-phase. The impact of common optical aberrations, such as sphericity, astigmatism, and coma, on the THz generation is also investigated. Aberrations are shown to generally decrease the laser-to-THz conversion efficiency and potentially induce spatial asymmetries and narrowing in the THz spectra.
... Water molecules are in translational and rotational diffusion motions. For example, the relaxation time of collective rotational diffusion ranges from subpicoseconds 10 13 s to ten picoseconds 10 11 s 19,20 . To understand the molecular hydration effect in detail, it is essential to measure the motion of water molecules in this time region. ...
... Spectroscopic measurements using THz light can measure molecular dynamics in the time domain of 10 12 s, that is picosecond range. It is known that there are four or five components in the absorption of water in the THz frequency region generally defined as the region from 0.1 to 10 THz 19,20 . From the lowest frequency, there is a collective rotational relaxation mode of water at about 20 GHz slow relaxation mode , a relaxation mode of isolated water molecules from the hydrogen bonding network at about 0.7 THz fast relaxation mode , a stretching vibration mode between water molecules at about 6 THz, and a libration mode between water molecules at about 15 THz. ...
Article
Full-text available
Phospholipids and surfactants form membranes and other self-assembled structures in water. However, it is not fully understood how the surrounding water (hydration water) is involved in their structure formation. In this paper, I summarize the results of our investigation of the long-range hydration state of phospholipids and surfactants at their surfaces by means of terahertz spectroscopy. By observing the collective rotational dynamics of water in the picosecond time scale, this technique allows us to observe not only the water directly bound to the solute, but also the weakly affected water outside of it. For example, PC phospholipids inhibit water dynamics over long distances, whereas PE phospholipids make water more mobile than bulk water. The causes of this difference in hydration and how it is involved in the structural formation of the membrane are reviewed. graphical abstract Fullsize Image
... This work focuses on one important substance in particular: liquid water and some of its diluted salt solutions. Water absorbs strongly at far-infrared [2][3][4][5][6][7][8][9] and terahertz [10][11][12][13][14][15][16][17][18][19][20][21] (THz) frequencies via the inter-molecular modes of hydrogen-bonded (HB) water molecules [22][23][24][25][26][27][28][29][30][31][32]. As an example, in Fig. 1 of Ref [33], the absorption coefficient of water roughly increases with THz frequency and reaches several thousand inverse centimeters, corresponding to penetration depths of a few microns. ...
Article
Full-text available
While bright terahertz sources are used to perform nonlinear experiments, they can be advantageous for high-precision linear measurements of opaque samples. By placing the sample away from the focus, nonlinearities can be suppressed, and sizeable amounts of transmitted radiation detected. Here, this approach is demonstrated for a 0.5 mm thick layer of liquid water in a static sample holder. Variations of the index of refraction as small as (7 ± 2) · 10⁻⁴ were detected at 0.58 THz for an aqueous salt solution containing ten millimoles of sodium chloride. To my knowledge, this precision is unprecedented in time-domain spectroscopy studies of diluted aqueous systems or other optically thick and opaque materials.
... These distinct characteristics enable the analysis of drug components, detection of counterfeit drugs, and screening for narcotics [4]. The THz waves uniquely correspond to the recombination and dissociation energies of water molecular clusters, endowing them with exceptional sensitivity to water [5,6]. This heightened sensitivity allows THz waves to precisely measure the distribution and quantity of water within the bio-samples, enabling their application in diagnostic devices for various diseases, such as tumors and burns [7,8,9,10,11]. ...
Article
Full-text available
Terahertz (THz) electromagnetic waves, known for their unique response to water, offer promising opportunities for next-generation biomedical diagnostics and novel cancer therapy technologies. This study investigated the impedance-matching effect, which enhances the efficiency of THz wave delivery into tissues and compensates for the signal distortion induced by the refractive index mismatch between the target and the sample substrate. Three candidate biocompatible materials, water, glycerol, and petroleum jelly were applied to a skin phantom and compared using THz two-dimensional imaging and time-of-flight imaging methods. Finally, we successfully demonstrated impedance-matching effect on mouse skin tissues.
... Indeed, the location and physical interpretation of the secondary relaxation mode has been the subject of ongoing debate [1]. The characteristic time τ 2 reported by earlier works ranges from 0.18 to 1.43 ps, and the amplitude is not negligible (1.60 < ∆ϵ 2 < 2.50) [1,49,[58][59][60][61][62]. Our simulation results suggest that the DPMD model again outperforms the SPC/Fw for predicting both the spectral amplitude and the characteristic time. ...
Preprint
Full-text available
Understanding the role of hydrogen bond networks in determining the relaxation dynamics is essential for understanding natural phenomena in liquid water. Classical pairwise additive models have been widely utilized for elaborating the underlying mechanism behind the relaxation phenomena. However, they have shown their limits due to either the absence or inaccurate descriptions of many-body and medium-to-long-range interactions. This work demonstrates that the Deep Potential Molecular Dynamics (DPMD) model help calculate the dielectric constant at the accuracy of the first-principles simulations. The DPMD model outperforms the classical force field (SPC/Fw) in predicting dielectric spectra especially in replicating high-frequency excesses, attributed to its adeptness in simulating intricate hydrogen bond networks. Through a comprehensive analysis of the simulation results, it becomes evident that only the DPMD model effectively accommodates a wide range of hydrogen bond coordination scenarios thereby characterizing the intricate nature of the hydrogen bond network. This adaptability stems from the intricate interplay of many-body interactions and intramolecular dynamics. In addition, orientation defects within the DPMD model play a significant role in shaping the potential energy barrier due to the adaptability.
Article
Performances of functional polymers are enhanced by designing the structures in the different hierarchies, such as monomer, polymer, secondary structures, and nanoscale morphologies. In the present work, an amorphous conjugated...
Article
Experiments investigating the properties of deeply supercooled liquid water are needed to develop a comprehensive understanding of water’s anomalous properties. One approach involves transiently heating nanoscale water films into the supercooled region for several nanoseconds at a time and then interrogating the water films after they have quenched to cryogenic temperatures. To relate the results obtained with this approach to other experiments and simulations on supercooled water, it is important to understand how closely the quenched structure tracks the (metastable) equilibrium structure of water as a function of the transient heating temperature. A key step involves quantifying the extent to which water that is transiently heated to ambient temperatures [hyperquenched water (HQW)] subsequently relaxes toward the structure of low-density amorphous (LDA) ice as it cools. We analyzed the infrared reflection–absorption spectra of LDA, HQW, and crystalline ice films to determine their complex indices of refraction. With this information, we estimate that HQW retains ∼50%–60% of a structural motif characteristic of water at high temperatures with the balance comprised of a low-temperature motif. This result, along with results from x-ray diffraction experiments on water and amorphous ices, allows one to quantify the fraction of the high-temperature motif at approximately zero pressure as a function of temperature from 150 to 350 K.
Article
Full-text available
That water can exist in two distinct `glassy' forms - low- and high-density amorphous ice - may provide the key to understanding some of the puzzling characteristics of cold and supercooled water, of which the glassy solids are more-viscous counterparts. Recent experimental and theoretical studies of both liquid and glassy water are now starting to offer the prospect of a coherent picture of the unusual properties of this ubiquitous substance.
Article
Full-text available
Although liquid water has been the focus of intensive research for over 100 years, a coherent physical picture that unifies all of the known anomalies of this liquid, is still lacking. Some of these anomalies occur in the supercooled region, and have been rationalized on the grounds of a possible retracing of the liquid-gas spinodal (metastability limit) line into the supercooled liquid region, or alternatively the presence of a line of first-order liquid-liquid phase transitions in this region which ends in a critical point,. But these ideas remain untested experimentally, in part because supercooled water can be probed only above the homogeneous nucleation temperature TH at which water spontaneously crystallizes. Here we report an experimental approach that is not restricted by the barrier imposed by TH, involving measurement of the decompression-induced melting curves of several high-pressure phases of ice in small emulsified droplets. We find that the melting curve for ice IV seems to undergo a discontinuity at precisely the location proposed for the line of liquid-liquid phase transitions. This is consistent with, but does not prove, the coexistence of two different phases of (supercooled) liquid water. From the experimental data we calculate a possible Gibbs potential surface and a corresponding equation of state for water, from the forms of which we estimate the coordinates of the liquid-liquid critical point to be at pressure Pc ~ 0.1GPa and temperature Tc ~ 220K.
Article
Full-text available
THE metastable extension of the phase diagram of liquid water exhibits rich features that manifest themselves in the equilibrium properties of water. For example, the density maximum at 4 °C and the minimum in the isothermal compressibility at 46 °C are thought to reflect the presence of singularities in the behaviour of thermodynamic quantities occurring in the supercooled region1 2. The 'stability-limit conjecture'3-5 suggests that these thermodynamic anomalies arise from a single limit of mechanical stability (spinodal line), originating at the liquid-gas critical point, which determines the limit of both superheating at high temperatures and supercooling at low temperatures. Here we present a comprehensive series of molecular dynamics simulations which suggest that, instead, the supercooling anomalies are caused by a newly identified critical point, above which the two metastable amorphous phases of ice (previously shown to be separated by a line of first-order transitions6,7) become indistinguishable. The two amorphous ice phases are thus incorporated into our understanding of the liquid state, providing a more complete picture of the metastable and stable behaviour of water.
Article
Full-text available
We evaluate thermodynamic, structural, and transport properties from extensive molecular-dynamics computer simulations of the ST2 and TIP4P models of liquid water over a wide range of thermodynamic states. We find a line in the phase diagram along which the isothermal compressibility of the supercooled liquid is a maximum. We further observe that along this line the magnitude of the maximum increases with decreasing temperature. Extrapolation to temperatures below those we are able to simulate suggests that the compressibility diverges. In this case, the line of compressibility maxima develops into a critical point followed at lower temperature by a line of first-order phase transitions. The behavior of structural and transport properties of simulated water supports the possibility of a line of first-order phase transitions separating two liquid phases differing in density. We therefore examine the experimentally known properties of liquid and amorphous solid water to test if the equation of state of the liquid might exhibit a line of compressibility maxima, possibly connected to a critical point that is the terminus of a line of phase transitions. We find that the currently available experimental data are consistent with these possibilities.
Article
Full-text available
We report measurements of the real and imaginary part of the dielectric constant of liquid water in the far-infrared region from 0.1 to 2.0 THz in a temperature range from 271.1 to 366.7 K. The data have been obtained with the use of THz time domain reflection spectroscopy, utilizing ultrashort electromagnetic pulses generated from a photoconductive antenna driven by femtosecond laser pulses. A Debye model with an additional relaxation time is used to fit the frequency dependence of the complex dielectric constants. We obtain a fast (fs) and a Debye (ps) relaxation time for the macroscopic polarization. The corresponding time correlation functions have been calculated with molecular dynamics simulations and are compared with experimental relaxation times. The temperature dependence of the Debye relaxation time is analyzed using three models: Transition state theory, a Debye–Stoke–Einstein relation between the viscosity and the Debye time, and a model stating that its temperature dependence can be extrapolated from a singularity of liquid water at 228 K. We find an excellent agreement between experiment and the two latter models. The simulations, however, present results with too large statistical error for establishing a relation for the temperature dependence. © 1997 American Institute of Physics.
Article
A femtosecond mid-infrared pump-probe study of the vibrational and orientational dynamics of the OH-stretching mode of HDO dissolved in D2O is presented. The orientational relaxation of the HDO molecules was observed to occur on either a very slow or a very fast time scale, with associated time constants of τR = 13 picoseconds and τR = 0.7 picosecond. It was observed that strongly hydrogen-bonded water molecules only relax through the slow orientational relaxation process, whereas the fast process dominates for weakly hydrogen-bonded molecules. This suggests that, with respect to orientional dynamics, two distinct molecular species exist in liquid water.
Article
We develop a new analytical equation of state for water based on the Song, Mason, and Ihm equation of state and Poole et al.’s simple model of the free energy of strong tetrahedral hydrogen bonds. Repulsive and attractive forces are modeled using a modification of the Weeks–Chandler–Anderson decomposition of the pair potential, with closed tetrahedral hydrogen bonds contributing both internal energy and entropy to the free energy of water. Strong tetrahedral hydrogen bonds are modeled explicitly using a simplified partition function. The resulting equation of state is 20–30 times more accurate than equivalent simple cubic equations of state over a wide range of pressures (0.1→3000 bar) and temperatures (−34→1200 °C) including the supercooled region. The new equation of state predicts a second liquid–liquid critical point at pC′ = 0.954 kbar, ρC′ = 1.045 g cm−3 and TC′ = 228.3 K. The temperature of this second critical point is above the homogeneous freezing temperature at 1 kbar, thus this region of the phase diagram may be experimentally accessible. The phase diagram also suggests that the homogeneous nucleation temperature above 1.2 kbar may be determined by a phase transition from high-density water to low-density water. © 1999 American Institute of Physics.