Vous êtes sur la page 1sur 5

Physics Letters A 375 (2011) 572576

Contents lists available at ScienceDirect

Physics Letters A
www.elsevier.com/locate/pla

Crossover between tetrahedral and hexagonal structures in liquid water


Osvaldo Chara 1 , Andrs N. McCarthy ,2 , J. Ral Grigera
Instituto de Fsica de Lquidos y Sistemas Biolgicos (IFLYSIB), CONICET - Universidad Nacional de La Plata, Argentina

a r t i c l e

i n f o

a b s t r a c t
It is widely accepted that liquid water structure is comprised of two closely interweaved components; i.e. tetrahedral (low density) and hexagonal (high density) structures. The relative amount of these components is temperature and pressure dependent. We propose an order parameter, based on the radial distribution function, that quanties the relative structural composition at any dened temperature and pressure, thus establishing the crossover point in structural dominance. At 300 K this point lies close to 2 kbar, pressure at which water looses most of its anomalous properties. 2010 Elsevier B.V. All rights reserved.

Article history: Received 18 July 2010 Received in revised form 24 November 2010 Accepted 25 November 2010 Available online 30 November 2010 Communicated by C.R. Doering Keywords: Water structure Two-state model Water radial distribution function components

1. Introduction The peculiar physical properties of water have for very long been recognized, being these particularities associated in general terms with its unique structure. From the seminal work of Bernal and Fowler [1] many suggestions on the structure of water and its behavior have been proposed [2]. Hall [3], based on ultrasonic experiments, supported the interpretation given by Bernal and Fowler of a two-state model, according to which low and high density structures are closely interweaved in water. Four decades later, further support for this model was given [47]. Subsequent studies have reinforced this concept, showing also its role on the phase transition of water [8]. Changes in temperature and pressure modify the properties of water. This can be interpreted as a consequence of changes in the ratio between high and low density structures, induced by the changes in thermodynamical conditions. The changes in water properties with temperature and pressure have been extensively shown through various experimental approaches [919]. Those results show that the so-called anomalies of water are gradually lost as temperature and pressure increase, moving towards a regime of a simple liquid. These changes establish that a region of intermediate behavior exists, for both increasing temperature and pressure.

Although temperature effects on the behavior of water have been extensively addressed [1619], studies of the effect of pressure on the dynamical and structural properties of water are much less abundant. Nevertheless, available data appears to be sucient to establish a pressure transition within the range of 1 to 3 kbar [20,21]. Considering water bulk in terms of the coexistence of two distinct structures (i.e. an open tetrahedral structure and a more compact hexagonal one), this transition may be interpreted in terms of the change in dominance of one structure over the other. Although this perspective has already been addressed, we still lack a robust quantitative parameter to characterize such a transition. In this work we present the development of an order parameter that denes the crossover between low and high density dominance in liquid water. We have used this parameter to analyze the radial distribution functions of water under various temperature and pressure conditions, using Argon as a reference for simple liquid behavior. The data analyzed was produced through Molecular Dynamic simulations. 2. Methods We carried out all Molecular Dynamic (MD) simulations using the GROMACS 4.0.4 package [22]. Water molecules were constrained using the SETTLE algorithm [23]. For the calculation of electrostatic forces we applied the Reaction Field method, with a cut-off radius of 1.4 nm, as well as for Lennard-Jones interactions. As starting congurations for water we used several replicas of 216 SPC/E water [24] molecule box equilibrated at 300 K and 1 bar. The system consisted of a cubic simulation box containing

Corresponding author. E-mail address: amccarthy@iysib.unlp.edu.ar (A.N. McCarthy). 1 Also at Center for Information Services and High Performance Computing, Dresden University of Technology, 01062 Dresden, Germany. 2 Also at Depto. Cs. Biolgicas, Fac. Cs. Exactas, Universidad Nacional de La Plata, Argentina. 0375-9601/$ see front matter 2010 Elsevier B.V. All rights reserved. doi:10.1016/j.physleta.2010.11.061

O. Chara et al. / Physics Letters A 375 (2011) 572576

573

In order to transform the above qualitative observations into quantitative structural information we use a two-state model to interpret each obtained g (r ). The model expresses the coexistence of two interweaved hydration contributions; i.e. an open tetrahedral structure and a more compact hexagonal one. Both contributions are treated qualitatively in the same way, taking into explicit account for each contribution the rst three hydration spheres, whilst the fourth sphere on is implicitly considered for both structures. Thus, each hydration contribution is modeled as a rst asymmetric distribution representing the rst hydration sphere, and two subsequent symmetric distributions representing the second and third hydration spheres. In both cases, the rst hydration sphere is modeled through a Freundlich distribution. For the tetrahedral and hexagonal contributions, the second and third hydration layers are modeled using a Gaussian distribution. Finally, all fourth and subsequent layers are modeled through a sigmoidal function. The sum of all explicit and implicit coordination contributions is what we call the G function, which may be explicitly dened as:
2 3

G (r ) =
j =1 i =2
Fig. 1. Evolution of g (r ) with pressure, at ambient temperature, as obtained through molecular dynamic simulations of the SPC/E water model. Inset: g (r ) of water at 1 bar (continuous curve), at 10 kbar (dotted curve) and liquid argon (dashed curve). Pressure appears to modify the g (r ) of water, resembling that of argon (see text main body).

f 1 (r ) + g i (r ) + S (r )

(1)

i where f j (r ) stands for the asymmetric Freundlich distribution,

a total of 1728 water molecules. All replicas were rst simulated during 2 ns, allowing for a complete equilibration of every pressure and temperature condition. Production data was collected during a complete nanosecond (1 ns) after equilibration. All replicas were weakly coupled to a thermal and hydrostatic bath, in order to work within the isothermalisobaric ensemble [25]. 3. Results and discussion In order to characterize the changes in water structure with pressure and temperature (and, hence, sensibly explain the experimentally observed changes in behavior), a number of molecular dynamic simulations were performed under various conditions, with pressure and temperature ranging from 1 bar to 10 kbar, and from 93 K to 373 K. Structure was evaluated through the oxygenoxygen radial distribution function, g (r ), which bears the advantage of having been amply studied through X-ray and neutron diffraction [12,17]. Fig. 1 shows the evolution of g (r ) with pressure, at ambient temperature, as obtained through molecular dynamic simulations of the SPC/E water model. In general terms, each g (r ) peak shows the preferential position of the successive neighboring water coordination spheres. The classical shape of the g (r ) of water at 1 bar may be observed both in the far plane of Fig. 1, and in its inset (black curve), highlighting that the position of the second peak typically reveals the tetrahedric component of the structure of water under these conditions. This contrasts with the position of the second peak of liquid argon g (r ), which reveals the presence of a simple (hexagonal) structure (Fig. 1 inset, gray curve). Nevertheless, as pressure increases, the second peak becomes attenuated, nally being replaced by a valley. Additionally, two further changes may be described: the appearance of a shoulder to the right of the rst peak and the closing in of the low pressure third peak, which nally results in the new second peak. Pressure appears to modify the g (r ) of water, resembling that of argon (Fig. 1 inset, light gray curve).

g ij (r ) stands for the symmetric Gaussian distribution, S (r ) stands for the coordination spheres implicitly accounted for, i subscript denotes the coordination sphere order, and j superscript denotes the tetrahedric or hexagonal contribution. The general explicit form of these functions is:

f 1 ( z ) = a j zb

j c j

with
2
i j2

z=r z = r0

r < 0

(2)

j g i (r )

ij /2
V rn

Ai

j (r )2

(3)

S (r ) =

Kn
j

+ rn
j j

(4)

where A i , i , i , a j , b j , c j , r0 , V , K and n are all the parameters involved. Fig. 2 shows the tetrahedral (black curves) and hexagonal (gray curves) components of the G function. Inset of Fig. 2 shows both the g (r ) obtained from simulation (dots) and the complete G function (dashdot curve). The correctness of the G function may be ultimately assessed through its ability to rebuild the g (r ) provided by molecular dynamic simulations. The G function was tted to the g (r ) data obtained for water at every temperature and pressure conditions simulated (for details of the tting algorithm see supplementary material). The same tting procedure was applied to the g (r ) data calculated for liquid Argon. For every simulated condition the tting procedure converged successfully, exemplied by Fig. 2 inset. The proposed model was thus validated, and consequently allows for its use in the quantication of the two coexisting structures, enabling the study of changes induced by pressure and temperature upon the relative contribution of these structures. To quantify these structure changes in uid, an order parameter is naturally desirable. Many parameters have consequently been proposed, the more relevant of which may be briey summarized as follows: The translational order parameter t , proposed by Truskett et al. [26], measures the tendency of pairs of molecules to adopt preferential separations, vanishing for an ideal gas, and adopting large

574

O. Chara et al. / Physics Letters A 375 (2011) 572576

Fig. 2. Complete G function and its components. The G function (dashdot black curve) tted to the g (r ) calculated from MD data of SPC/E water model. Tetrahet dral structural contributions are represented by f 1 (rst coordination sphere, dotted t t black curve), g 2 (second coordination sphere, continuous black curve) and g 3 (third coordination sphere, dashed black curve). Hexagonal structural contributions are h h represented by f 1 (rst coordination sphere, dotted grey curve), g 2 (second coorh (third coordination sphere, dashed dination sphere, continuous grey curve) and g 3 grey curve). S represents the fourth sphere on, for both structural contributions (continuous light gray curve). Inset: g (r ) obtained from MD simulation (dots) and the G function (dashdot curve).

Fig. 3. The dimensionless parameter P r of SPC/E water model at 300 K as a function of pressure ( P , kbar). The crossover point is dened by the pressure at which P r = 0.

values for a crystal. This parameter gives an idea of structure in a general sense. Nevertheless, in order to discriminate the nature of the structure, the use of other parameters is required. Steinhardt et al. [27] dened the Q 3 parameter whilst Naberukhin et al. [28,29] proposed the use of T parameter, both being capable of testing tetrahedrality. As discussed by Chau and Hardwick (1998) [30], Q 3 suffers from two drawbacks. First, one can obtain the tetrahedral value for Q 3 only when the conguration is of a certain pyramidal shape. Second, its value for a tetrahedral conguration is not a limiting value, which in some cases could result in a mean value near tetrahedrality, although the congurations would be far from tetrahedral. The other parameter, T [28,29], also has two main drawbacks: given that each evaluation requires the calculation of ten distances it is expensive in computer time; likewise, it is dicult to connect its value to the congurational geometry. In order to solve the aforementioned drawbacks, Chau and Hardwick [30] proposed the parameter S , which is composed of two parts, S g and S k . The rst part, S g , is used to test for tetrahedral angles, and its value can be mapped to the congurational geometry. The second part, S k , is the variance in the radial distance from the central vertex to the other four vertices. The parameter S has the advantage that it can tell us how distorted the polyhedron is, and in what way. A drawback of S is that it is dicult to connect the value of S k to the geometry of the conguration. In an extensive and detailed study, Errington and Debenedetti [31] used a rescaled version of the parameter S (which they call q ) to study the changes in structure of the SPC/E water model under various temperature and pressure conditions. Here the authors show a bimodal distribution for this order parameter, suggesting that each water molecule and its four nearest neighbors present a transient arrangement [3234] which can be described as being predominantly structured (and ice-like) or unstructured. All the aforementioned parameters are constructed upon the knowledge of the position of every oxygen atom in water bulk. This can only be achieved through computational studies of water models; i.e. MD or Monte Carlo simulations. Additionally, all the

discussed parameters that evaluate tetrahedrality only take explicitly into account the rst coordination sphere. We propose a new order parameter ( P r ) which is capable of evaluating structure beyond the rst coordination sphere. Furthermore, this parameter requires only of the knowledge of the probability density function for the position of water oxygens. Such information may be obtained by both computational and experimental studies. The parameter P r is dened as:

Pr =
with

Ct Ch Ct + Ch At 1 wt 1 Ah 1 wh 1
j

(5)

Ct = Ch =

+ +

At 2 wt 2 Ah 2 wh 2

(6)

(7)

where A i is proportional to the area under the Gaussian component i of the tetrahedral ( j = t ) and hexagonal ( j = h ) contribution j respectively, and w i gives a measure of the dispersion of the corresponding Gaussian component. Fig. 3 shows the P r at 300 K as a function of pressure. A monotonically decreasing behavior from 1 bar to 10 kbar may be apparently appreciated. The actual limiting values of P r go from 0.2 at 1 bar to 0.38 at 10 kbar. The same analysis was applied to the radial distribution functions of liquid argon produced from a number of molecular dynamic simulations performed under various pressure conditions, ranging from 60 bar to 180 bar, obtaining values of P r within the range of 0.43 to 0.46. This is close to the limiting value found for water. Hence, although from a mathematical point of view 1 > P r > 1, the extreme values (1 and 1) are found only in pure tetrahedral and hexagonal structures; i.e. in crystals. In opposition, the liquid state may be regarded as a closely interweaved coexistence of different structures, eventually showing a dominance of one over the rest. The value of P r becomes negative at pressures above 2.3 kbar. This point marks the crossover between the prevalence of tetrahedral and hexagonal structures. The calculated crossover point agrees fairly well with the transition of water properties observed experimentally, such as rota-

O. Chara et al. / Physics Letters A 375 (2011) 572576

575

changes are observed in both structural and dynamical properties of water. From the previous results, the critical point may be estimated at T c = 268 K and P c = 3.5 kbar. This value is very much within the order of that obtained for the TIP5P ( T c = 217 K; P c = 3.5 kbar) and ST2 ( T c = 246 K; P c = 1.86 kbar) water models, which were calculated using the changes in compressibility and diffusion coefcient [38]. This result is likewise consistent with those obtained by Kumar et al. [40]. The behavior of these two properties come within the so-called water anomalies. 4. Conclusions Based on a two-state model for liquid water, we have explicitly taken into account the rst three coordination spheres for both the tetrahedral and hexagonal structural contributions. We thus elaborated the G function in order to reproduce the oxygenoxygen probability density function. This function has shown a remarkable ability to recreate this probability function at every temperature and pressure condition tested. The data provided by our model was used for the elaboration of a structural order parameter ( P r ) that allows us to dene a structural crossover point which establishes the change in prevalence between the tetrahedral and hexagonal structural contributions with pressure and temperature. The proposed parameter has two further advantages. Namely, that it is able to quantify structural contributions beyond the rst coordination sphere and that it may be applied to the study of computationally or experimentally obtained gr functions. We may conclude that the increase in both pressure and temperature induce an increase in the hexagonal structural contribution in detriment of the tetrahedral one. Acknowledgements The authors thank S.A. Grigera and L.A. Pugnaloni for useful discussions. This work was partially supported by the National Research Council of Argentina (CONICET, PIP-112-200801-03247), the National University of La Plata and the National Agency for the Promotion of Science and Technology (ANPCyT, PICT-1564). O.C. and J.R.G. are members of the Research Career of the CONICET, Argentina. A.N.McC. is member of the Research Career of the Research Council of the Buenos Aires Province (CICPBA), Argentina. References
[1] J.D. Bernal, R.H. Fowler, J. Chem. Phys. 1 (1933) 515. [2] H.J.C. Berendsen, in: E. Cole (Ed.), Water Structure in Theoretical and Experimental Biophysics, Marcel Dekker, 1967 (Chapter 1). [3] L. Hall, Phys. Rev. 73 (1948) 775. [4] A. Geiger, H.E. Stanley, Phys. Rev. Lett. 49 (1982) 1749. [5] C.M. Davis, T.A. Litovitz, J. Chem. Phys. 42 (1990) 2563. [6] C.M. Davis, D. Bradley, J. Chem. Phys. 45 (1990) 2461. [7] M.C. Bellissent-Funel, J. Teixeira, L.J. Bosio, J. Chem. Phys. 87 (1990) 2231. [8] O. Mishima, H. Stanley, Nature 396 (1998) 329. [9] E.W. Castner, Y.J. Chang, Y.C. Chu, G.E. Walrafen, J. Chem. Phys. 102 (1995) 653. [10] C.J. Montrose, J.A. Bucaro, J. Marshall-Coakley, T.A. Litovitz, J. Chem. Phys. 60 (1974) 5025. [11] W. Danninger, G. Zundel, J. Chem. Phys. 74 (1981) 2779. [12] J. Yeixeira, S.H. Chen, M.-C. Bellissent-Funel, J. Mol. Liq. 48 (1991) 111. [13] R. Laenen, C. Rauscher, A. Laubereau, Phys. Rev. Lett. 80 (1998) 2622. [14] S. Woutersen, U. Emmerichs, H.J. Bakker, Science 278 (1997) 658. [15] H.K. Nienhuys, S. Woutersen, R.A. van Santen, H.J. Bakker, J. Chem. Phys. 111 (1999) 1494. [16] G.M. Gale, G. Gallot, F. Hache, N. Lascoux, S. Bratos, J.C. Leicknam, Phys. Rev. Lett. 82 (1999) 1068. [17] A.H. Narten, M.D. Danford, H.A. Levy, Faraday Discuss. 43 (1967) 97. [18] A.K. Soper, F. Bruni, M.A. Ricci, J. Chem. Phys. 106 (1997) 247. [19] K. Modig, B.G. Pfrommer, B. Halle, Phys. Rev. Lett. 90 (2003) 075502. [20] F.X. Prielmeier, E.W. Lang, R.J. Speedy, H.D. Ldemann, Phys. Rev. Lett. 59 (1987) 1128.

Fig. 4. Dependence of the crossover point with temperature (closed black circles). The points under 243 K belong to the supercooled region of the SPC/E water model. HDL shows the region in which the prevalent structure corresponds to the hexagonal coordination state, which is structurally compatible with a high density condition of water. The region that corresponds to the predominance of the tetrahedric structure has been denoted as LDL, and is structurally compatible with the lower density condition of water. The open circle indicates the critical point. The Widom line can be observed for temperatures above the critical point (black line). The coexistence line is represented by the broken line. The crossover points at temperatures below the critical point were tted with a non-linear curve (black non-linear curve).

tional and translational diffusion [20,35], viscosity [21], and inelastic X-ray diffraction [36,37]. In order to analyze the temperature dependence of the crossover point, further simulations were performed at temperatures ranging from 93 K to 373 K. Within the region comprised between 243 K and 373 K, a monotonically decreasing behavior of this parameter with the increase in temperature becomes apparent. Such an observation comes as a reasonable expectation when we take into consideration that temperature holds a destructive effect upon the tetrahedral structure of water. These results, together with this last interpretation, are in good agreement with the changes in water structure observed by Errington and Debenedetti [31]. Additionally, we have calculated the crossover points for different temperatures, including points in the supercooled range. The results are shown in Fig. 4. The calculated points dene a crossover curve separating two regions. The region in which the prevalent structure corresponds to a hexagonal coordination has been indicated as HDL (high density liquid [38]), which is structurally compatible with the high density condition of water, whilst the denomination of LDL (low density liquid [38]) was used for the region that corresponds to the predominance of tetrahedric structure, which in turn is structurally compatible with the lower density condition of water. The results obtained agree qualitatively with such a behavior as expected for supercooled water [38,39]. The obtained crossover data shows a critical point ( T c , P c ) between a linear and a non-linear regime. The linear regime corresponds to the so-called Widom line [38], which at lower temperatures can be extrapolated as a coexistence region (broken line in Fig. 4). The line connecting the different points can be regarded as a frontier that delimitates structural prevalence (full curve). When crossing the Widom line along an isobar at P < P c (path a), a transition from the HDL to LDL is obtained, resulting in qualitative changes in behavior of structural and dynamical properties. However, when an isobar with P > P c (path b) is followed, continuous

576

O. Chara et al. / Physics Letters A 375 (2011) 572576

[21] R.A. Horne, D.S. Johnson, J. Phys. Chem. 70 (1966) 2182. [22] D. van der Spoel, E. Lindahl, B. Hess, G. Groenhof, A.E. Mark, H.J.C. Berendsen, J. Comput. Chem. 26 (2005) 1701. [23] S. Miyamoto, P.A. Kollman, J. Comput. Chem. 13 (1992) 952. [24] H.J.C. Berendsen, J.R. Grigera, T.P. Straatsma, J. Phys. Chem. 91 (1987) 6269. [25] H.J.C. Berendsen, J.P.M. Postma, A. DiNola, J.R. Haak, J. Chem. Phys. 81 (1984) 3684. [26] T.M. Truskett, S. Torquato, P.G. Debenedetti, Phys. Rev. E 62 (2000) 993. [27] P.J. Steinhardt, D.R. Nelson, M. Ronchetti, Phys. Rev. B 28 (1983) 784. [28] Y.U.I. Naberukhin, V.P. Voloshin, N.N. Medvedev, Molec. Phys. 73 (1991) 917. [29] M. Kiselev, M. Poxleitner, J. Seitz-Beywl, K. Heinzinger, Z. Naturf. A 48 (1993) 806.

[30] [31] [32] [33] [34] [35] [36] [37]

P.L. Chau, A.J. Hardwick, Molec. Phys. 93 (1998) 511. J.R. Errington, P.G. Debenedetti, Nature 409 (2001) 318. E. Shiratani, M. Sasai, J. Chem. Phys. 108 (1998) 3264. F. Sciortino, P.H. Poole, H.E. Stanley, S. Havlin, Phys. Rev. Lett. 64 (1990) 1686. F. Sciortino, A. Geiger, H.E. Stanley, Nature 354 (1991) 218. L.A. Wolf, J. Chem. Soc., Faraday Trans. 71 (1975) 784. Y.Q. Cai, et al., Phys. Rev. Lett. 94 (2005) 025502. A. Cunsolo, F. Formisano, C. Ferrero, F. Bencivenga, S. Finet, J. Chem. Phys. 131 (2009) 194502. [38] L. Xu, P. Kumar, S. V Buldyrev, S.H. Chen, P.H. Poole, F. Sciortino, H.E. Stanley, PNAS 102 (2005) 16562. [39] G. Franzese1, H. Stanley, J. Phys.: Condens. Matter 19 (2007) 205126. [40] P. Kumar, S.V. Buldyrev, S.R. Becker, P.H. Poole, F.W. Starr, H.E. Stanley, PNAS 105 (2007) 9575.

Vous aimerez peut-être aussi