transition from one-dimensional water to ferroelectric ice ...bonded to one o3w and one o2w) form a...

6
Transition from one-dimensional water to ferroelectric ice within a supramolecular architecture Hai-Xia Zhao a,1 , Xiang-Jian Kong a,1 , Hui Li b,1 , Yi-Chang Jin a , La-Sheng Long a,2 , Xiao Cheng Zeng b,2 , Rong-Bin Huang a , and Lan-Sun Zheng a a State Key Laboratory of Physical Chemistry of Solid Surface and Department of Chemistry, College of Chemistry and Chemical Engineering, Xiamen University, Xiamen 361005, China; and b Department of Chemistry, University of Nebraska, Lincoln, NE 68588 Edited* by H. Eugene Stanley, Boston University, Boston, MA, and approved January 12, 2011 (received for review July 16, 2010) Ferroelectric materials are characterized by spontaneous electric polarization that can be reversed by inverting an external electric field. Owing to their unique properties, ferroelectric materials have found broad applications in microelectronics, computers, and trans- ducers. Water molecules are dipolar and thus ferroelectric alignment of water molecules is conceivable when water freezes into special forms of ice. Although the ferroelectric ice XI has been proposed to exist on Uranus, Neptune, or Pluto, evidence of a fully proton- ordered ferroelectric ice is still elusive. To date, existence of ferroelec- tric ice with partial ferroelectric alignment has been demonstrated only in thin films of ice grown on platinum surfaces or within micro- domains of alkali-hydroxide doped ice I. Here we report a unique structure of quasi-one-dimensional ðH 2 OÞ 12n wire confined to a 3D supramolecular architecture of ½Cu I 2 Cu II ðCDTAÞð4,4 0 -bpyÞ 2 n H 4 CDTA, trans-1,2-diaminocyclohexane-N,N,N,N-tetraacetic acid; 4,4-bpy, 4,4-bipyridine). In stark contrast to the bulk, this 1D water wire not only exhibits enormous dielectric anomalies at approximately 175 and 277 K, respectively, but also undergoes a spontaneous transi- tion between 1D liquidand 1D ferroelectric iceat approximately 277 K. Hitherto unrevealed properties of the 1D water wire will be valuable to the understanding of anomalous properties of water and synthesis of novel ferroelectric materials. ab initio molecular dynamics phase transition supramolecular nanochannel A n understanding of the properties of water wire and quasi-1D ice confined to a nanoscale channel has important implica- tions for the biological sciences, geoscience, and nanoscience (15). The nanoscale channel can endow the water wire and 1D ice with unusual properties that differ from 3D water and ice. Hence, insights into these properties on the molecular level not only can offer explanation of intriguing phenomena such as pro- ton and water transport in cell membranes (2) and the formation of a 1D ice tube with a water chain at the center (5), but can also assist synthesis of new nanochannels to influence the properties of water wire and 1D ice by changing the environment. In this article, our attention will be placed on the spontaneous formation of a 1D ferroelectric ice in a newly synthesized nanochannel. Does standalone single-phase ferroelectric ice exist in nature? This question has fascinated scientists for decades (614). If the single-phase ferroelectric ice does exist, it would have a net polarization toward one direction (15) because of the ordered alignment of the water molecules. However, achieving a standa- lone and pure 3D ferroelectric ice seems to be a formidable task in the laboratory because the full phase-transformation time required for a bulk 3D ferroelectric ice is estimated to be on the order of 10 5 y (8, 9) without any assistance of dopants as cat- alysts. To date, all ferroelectric ices produced in the laboratory are low dimensional and in mixed phases (1012). Iedema et al. (10) revealed amorphous or cubic ice films, deposited on a pla- tinum surface, in which only about 0.2% of molecules exhibits ferroelectric alignment. Su et al. demonstrated a nanofilm of hex- agonal ice, also deposited on a platinum surface, which shows a net polarization decaying with the distance from the platinum surface (11). Even protein hydration shells can be polarized into a ferroelectric cluster with a large averaged dipole moment (16). Indeed, in a low-dimensional environment, either on a substrate or in a nanochannel, the water-surface interaction may effectively hinder some protons from going to the disordered state (espe- cially at low temperature) (1719). Hence, well-designed nano- channels have been viewed as a promising host to realizing low-dimensional single-phase ferroelectric ice (13, 14). For exam- ple, five distinctive phases of quasi-1D ice, formed inside carbon nanotubes, have been revealed experimentally (2023). These five 1D ice phases all exhibit single-walled tubular morphologies, including the pentagon ice nanotube (24) (provisionally named 1D ice I; ref. 18), hexagon ice nanotube (1D ice II), heptagon ice nanotube (1D ice III), octagon ice nanotube (1D ice IV), and nonagon ice nanotube (1D ice V). Additionally, a core-sheath 1D ice structure has been detected in a neutron scattering experi- ment (5) where the core is a single-file water chain and the sheath is an octagon ice nanotube (provisionally named as 1D ice IVa). Previous classical molecular dynamics simulations (13, 14) have suggested that the pentagon and heptagon ice nanotubes are 1D ice with mixed ferroelectric/antiferroelectric domains. Besides the carbon nanotubes, a large number of organic and inorganic species have been successfully synthesized as hosts for confining water on the nanoscale (4, 17, 2528). Nevertheless, spe- cies that can lead to single-phase ferroelectric ice have not been demonstrated experimentally. Recently, we reported that a water wire confined to the nanochannel ½La 2 Cu 3 fNHðCH 2 COOÞ 2 g 6 n undergoes a transition between 1D liquid and 1D antiferroelectric ice (provisionally named as 1D ice VI) at approximately 350 K (17). This finding indicates that porous metal-organic frameworks not only provide a tubular host system to understand anomalous properties of water, but also may be exploited to achieve single- phase 1D ferroelectric ice. Herein, we report the unique structure and unusual properties of 1D ½ðH 2 OÞ 12 n wire confined to a newly synthesized nanochannel, namely, 3D supramolecular architecture of ½Cu I 2 Cu II ðCDTAÞð4;4 0 -bpyÞ 2 n [H 4 CDTA, trans-1,2-diamino- cyclohexane-N,N,N,N-tetraacetic acid; 4,4-bpy, 4,4-bipyridine). Results and Discussion Our single-crystal structural analysis shows that the structure of the 1D ðH 2 OÞ 12n wire has a periodic 12-mer framework held Author contributions: L.-S.L. and X.C.Z. designed research; H.-X.Z., X.-J.K., H.L., Y.-C.J., L.-S.L., and X.C.Z. performed research; L.-S.L., X.C.Z., R.-B.H., and L.-S.Z. contributed new reagents/analytic tools; H.-X.Z., X.-J.K., H.L., Y.-C.J., L.-S.L., X.C.Z., R.-B.H., and L.-S.Z. analyzed data; and L.-S.L., X.C.Z., and L.-S.Z. wrote the paper. The authors declare no conflict of interest. *This Direct Submission article had a prearranged editor. Data deposition: The atomic coordinates have been deposited in the Cambridge Structural Database, Cambridge Crystallographic Data Centre, Cambridge CB2 1EZ, United Kingdom, www.ccdc.cam.ac.uk/data_request/cif (CSD reference nos. 800743800748). 1 H.-X.Z., X.-J.K., and H.L. contributed equally to this work. 2 To whom correspondence may be addressed. E-mail: [email protected] or xzeng1@ unl.edu. This article contains supporting information online at www.pnas.org/lookup/suppl/ doi:10.1073/pnas.1010310108/-/DCSupplemental. www.pnas.org/cgi/doi/10.1073/pnas.1010310108 PNAS March 1, 2011 vol. 108 no. 9 34813486 APPLIED PHYSICAL SCIENCES Downloaded by guest on May 25, 2021

Upload: others

Post on 22-Jan-2021

0 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Transition from one-dimensional water to ferroelectric ice ...bonded to one O3w and one O2w) form a cyclic and boat-like water octamer. Each crystallographically independent water

Transition from one-dimensional water to ferroelectricice within a supramolecular architectureHai-Xia Zhaoa,1, Xiang-Jian Konga,1, Hui Lib,1, Yi-Chang Jina, La-Sheng Longa,2, Xiao Cheng Zengb,2,Rong-Bin Huanga, and Lan-Sun Zhenga

aState Key Laboratory of Physical Chemistry of Solid Surface and Department of Chemistry, College of Chemistry and Chemical Engineering,Xiamen University, Xiamen 361005, China; and bDepartment of Chemistry, University of Nebraska, Lincoln, NE 68588

Edited* by H. Eugene Stanley, Boston University, Boston, MA, and approved January 12, 2011 (received for review July 16, 2010)

Ferroelectric materials are characterized by spontaneous electricpolarization that can be reversed by inverting an external electricfield. Owing to their unique properties, ferroelectric materials havefound broad applications in microelectronics, computers, and trans-ducers. Water molecules are dipolar and thus ferroelectric alignmentof water molecules is conceivable when water freezes into specialforms of ice. Although the ferroelectric ice XI has been proposedto exist on Uranus, Neptune, or Pluto, evidence of a fully proton-ordered ferroelectric ice is still elusive. To date, existence of ferroelec-tric ice with partial ferroelectric alignment has been demonstratedonly in thin films of ice grown on platinum surfaces or within micro-domains of alkali-hydroxide doped ice I. Here we report a uniquestructure of quasi-one-dimensional ðH2OÞ12n wire confined to a 3Dsupramolecular architecture of ½CuI

2CuIIðCDTAÞð4,40-bpyÞ2�n H4CDTA,

trans-1,2-diaminocyclohexane-N,N,N′,N′-tetraacetic acid; 4,4′-bpy,4,4′-bipyridine). In stark contrast to the bulk, this 1D water wire notonly exhibits enormous dielectric anomalies at approximately 175and 277 K, respectively, but also undergoes a spontaneous transi-tion between “1D liquid” and “1D ferroelectric ice” at approximately277 K. Hitherto unrevealed properties of the 1D water wire will bevaluable to the understanding of anomalous properties ofwater andsynthesis of novel ferroelectric materials.

ab initio molecular dynamics ∣ phase transition ∣supramolecular nanochannel

An understanding of the properties of water wire and quasi-1Dice confined to a nanoscale channel has important implica-

tions for the biological sciences, geoscience, and nanoscience(1–5). The nanoscale channel can endow the water wire and 1Dice with unusual properties that differ from 3D water and ice.Hence, insights into these properties on the molecular level notonly can offer explanation of intriguing phenomena such as pro-ton and water transport in cell membranes (2) and the formationof a 1D ice tube with a water chain at the center (5), but can alsoassist synthesis of new nanochannels to influence the propertiesof water wire and 1D ice by changing the environment. In thisarticle, our attention will be placed on the spontaneous formationof a 1D ferroelectric ice in a newly synthesized nanochannel.

Does standalone single-phase ferroelectric ice exist in nature?This question has fascinated scientists for decades (6–14). If thesingle-phase ferroelectric ice does exist, it would have a netpolarization toward one direction (15) because of the orderedalignment of the water molecules. However, achieving a standa-lone and pure 3D ferroelectric ice seems to be a formidable taskin the laboratory because the full phase-transformation timerequired for a bulk 3D ferroelectric ice is estimated to be onthe order of 105 y (8, 9) without any assistance of dopants as cat-alysts. To date, all ferroelectric ices produced in the laboratoryare low dimensional and in mixed phases (10–12). Iedema et al.(10) revealed amorphous or cubic ice films, deposited on a pla-tinum surface, in which only about 0.2% of molecules exhibitsferroelectric alignment. Su et al. demonstrated a nanofilm of hex-agonal ice, also deposited on a platinum surface, which shows anet polarization decaying with the distance from the platinum

surface (11). Even protein hydration shells can be polarized intoa ferroelectric cluster with a large averaged dipole moment (16).Indeed, in a low-dimensional environment, either on a substrateor in a nanochannel, the water-surface interaction may effectivelyhinder some protons from going to the disordered state (espe-cially at low temperature) (17–19). Hence, well-designed nano-channels have been viewed as a promising host to realizinglow-dimensional single-phase ferroelectric ice (13, 14). For exam-ple, five distinctive phases of quasi-1D ice, formed inside carbonnanotubes, have been revealed experimentally (20–23). Thesefive 1D ice phases all exhibit single-walled tubular morphologies,including the pentagon ice nanotube (24) (provisionally named1D ice I; ref. 18), hexagon ice nanotube (1D ice II), heptagon icenanotube (1D ice III), octagon ice nanotube (1D ice IV), andnonagon ice nanotube (1D ice V). Additionally, a core-sheath1D ice structure has been detected in a neutron scattering experi-ment (5) where the core is a single-file water chain and the sheathis an octagon ice nanotube (provisionally named as 1D ice IVa).Previous classical molecular dynamics simulations (13, 14) havesuggested that the pentagon and heptagon ice nanotubes are1D ice with mixed ferroelectric/antiferroelectric domains.

Besides the carbon nanotubes, a large number of organic andinorganic species have been successfully synthesized as hosts forconfining water on the nanoscale (4, 17, 25–28). Nevertheless, spe-cies that can lead to single-phase ferroelectric ice have not beendemonstrated experimentally. Recently, we reported that a waterwire confined to the nanochannel ½La2Cu3fNHðCH2COOÞ2g6�nundergoes a transition between 1D liquid and 1D antiferroelectricice (provisionally named as 1D ice VI) at approximately 350 K(17). This finding indicates that porous metal-organic frameworksnot only provide a tubular host system to understand anomalousproperties of water, but also may be exploited to achieve single-phase 1D ferroelectric ice. Herein, we report the unique structureand unusual properties of 1D ½ðH2OÞ12�n wire confined to a newlysynthesized nanochannel, namely, 3D supramolecular architectureof ½CuI2CuIIðCDTAÞð4;40-bpyÞ2�n [H4CDTA, trans-1,2-diamino-cyclohexane-N,N,N′,N′-tetraacetic acid; 4,4′-bpy, 4,4′-bipyridine).

Results and DiscussionOur single-crystal structural analysis shows that the structure ofthe 1D ðH2OÞ12n wire has a periodic 12-mer framework held

Author contributions: L.-S.L. and X.C.Z. designed research; H.-X.Z., X.-J.K., H.L., Y.-C.J.,L.-S.L., and X.C.Z. performed research; L.-S.L., X.C.Z., R.-B.H., and L.-S.Z. contributed newreagents/analytic tools; H.-X.Z., X.-J.K., H.L., Y.-C.J., L.-S.L., X.C.Z., R.-B.H., and L.-S.Z.analyzed data; and L.-S.L., X.C.Z., and L.-S.Z. wrote the paper.

The authors declare no conflict of interest.

*This Direct Submission article had a prearranged editor.

Data deposition: The atomic coordinates have been deposited in the CambridgeStructural Database, Cambridge Crystallographic Data Centre, Cambridge CB2 1EZ, UnitedKingdom, www.ccdc.cam.ac.uk/data_request/cif (CSD reference nos. 800743–800748).1H.-X.Z., X.-J.K., and H.L. contributed equally to this work.2To whom correspondence may be addressed. E-mail: [email protected] or [email protected].

This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.1073/pnas.1010310108/-/DCSupplemental.

www.pnas.org/cgi/doi/10.1073/pnas.1010310108 PNAS ∣ March 1, 2011 ∣ vol. 108 ∣ no. 9 ∣ 3481–3486

APP

LIED

PHYS

ICAL

SCIENCE

S

Dow

nloa

ded

by g

uest

on

May

25,

202

1

Page 2: Transition from one-dimensional water to ferroelectric ice ...bonded to one O3w and one O2w) form a cyclic and boat-like water octamer. Each crystallographically independent water

together by hydrogen-bonding interactions, i.e., each unit cellcontains 12 water molecules. The geometry of the 12 oxygenatoms in a unit cell can be described as a six-rung ladder withone rung at one end and two rungs at another end being rotatedby approximately 90° with respective to the central axis of thenanochannel, while the oxygen ladder itself is along the axialdirection of the nanochannel. Measurement of the temperature-dependent dielectric constant of the water wire reveals that thewater wire exhibits large dielectric anomalies at approximately175 and 277 K, respectively, significantly different from that ofbulk water. More interestingly, the measurement of temperature-and frequency-dependent dielectric constants and temperature-dependent dielectric hysteresis of the 1D water wire indicatesthat the 1D wire undergoes a phase transition between 1D ferro-electric ice and 1D liquid at approximately 277 K. A single-phaseferroelectric ice is now produced within a nanochannel.

This 1D ice is spontaneously formed within a 3D supramolecu-lar architecture of ½CuI2CuIIðCDTAÞð4;40-bpyÞ2�n • 6nðH2OÞ (1)(Dataset S1) prepared by the reaction of H4CDTA (0.364 g,1.0 mmol), CuðCH3COOÞ2 •H2O (0.199 g, 1.0 mmol), 4,4′-bpy(0.156 g, 1.0 mmol), and 10 mL water under the hydrothermalcondition. Crystallographic analysis at 123 K indicates that theasymmetric unit of (1) contains half an independent divalentCuII ion, one monovalent CuI, half a CDTA4− ligand, one4,4′-bpy ligand, and three water molecules. Each CuII ion, locatedin the octahedral structure, is coordinated by two N atoms andfour modentate carboxylate groups from one CDTA4− ligandto form a metalloligand ½CuIICDTA�2−. Each CuI ion, locatedin the T-shaped structure, is coordinated by two N atoms, respec-tively, from two 4,4′-bpy ligands and one monodentate carboxy-

late group from one metalloligand ½CuIICDTA�2− (Fig. 1A). TheT-shaped coordination mode for the copper ion indicates exis-tence of the CuI species, consistent with the observation of mag-netic property of (1) (Fig. S1).

The two adjacent CuI ions, bridged by one 4,4′-bpy ligand, inwhich its terminal N atom coordinates with one CuI ion, formsa 1D chain of ½CuIð4;40-bpyÞ�nnþ. Connection of the adjacent1D chains of ½CuIð4;40-bpyÞ�nnþ through the metalloligand of½CuIICDTA�2− with its two O atoms coordinating respectivelywith two CuI ions from two adjacent 1D chains leads to a 2D 63

network ½CuI2CuIIðCDTAÞð4;40-bpyÞ2�n (Fig. 1B) which furtherextends to a threefold interpenetration network as shown inFig. 1 C and D. The 3D supramolecular architecture of (1)can be viewed such that the 1D water wire ðH2OÞ12n acts asproton donor and is hydrogen bonded to the carboxylate of½CuIICDTA�2− from the adjacent threefold interpenetration net-works, as illustrated in Fig. 1E.

In the 1D ice ðH2OÞ12n, there exist three crystallographicallyindependent water molecules, O1w, O2w, and O3w (Fig. 1F).Each O1w acts as both proton accepter and donor to connect withthe adjacent two O1w, forming a cyclic water tetramer whoseplane is slightly twisted but normal to the central axis of thenanochannel. Four O2w (each hydrogen-bonded to one O2wand one O3w, respectively) and four O3w (each hydrogen-bonded to one O3w and one O2w) form a cyclic and boat-likewater octamer. Each crystallographically independent watermolecule has a fully occupied H atom (H1wa for O1w, H2wafor O2w, and H3wa for O3w). The fully occupied H atoms forO1w and O2w are hydrogen-bonded, respectively, to O4 and O1from carboxylate of ½CuIICDTA�2−, whereas O3w is hydrogen-

Fig. 1. (A) Ball and stick view of the coordination environment of CuI and CuII ions. (B) The 2D network ½CuICuIIðCDTAÞðbpyÞ2�n viewed along the c axis.(C) The threefold interpenetration network of ½CuICuIIðCDTAÞðbpyÞ2�n viewed along the c axis. (D) The threefold interpenetration network viewed alongthe b axis. (E) The 3D supramolecular architecture of (1) along the a axis. (F) A side view of water structures in two unit cells of the 1D water wireðH2OÞ12n in (1). [Note: For the three crystallographically independent water molecules in the 1D ice, they are plotted in such a way that each water moleculeseems to have three hydrogen atoms, although in reality each water molecule only has two hydrogen atoms. This way of plotting is commonly used (e.g., inref. 4) to reflect that one of the two hydrogen atoms in every water molecule is positionally disordered and thus is treated as half-occupancy as pointed out inthe text. From the viewpoint of crystallography, these dynamic hydrogen atoms are alternatively hydrogen-bonded to adjacent water molecules in the 1D ice.]Red and green balls denote oxygen atoms in the front and the back positions, respectively. The blue dashed lines denote hydrogen bonds. O1–O4 are oxygenatoms of the nanochannel.

3482 ∣ www.pnas.org/cgi/doi/10.1073/pnas.1010310108 Zhao et al.

Dow

nloa

ded

by g

uest

on

May

25,

202

1

Page 3: Transition from one-dimensional water to ferroelectric ice ...bonded to one O3w and one O2w) form a cyclic and boat-like water octamer. Each crystallographically independent water

bonded to O1w. The remaining H atoms of the water moleculesare disordered and treated as half-occupied. As a consequence,the 1D ice ðH2OÞ12n can be viewed as a water tetramer acting as aproton acceptor and boat-like water octamer acting as a protondonor (Fig. 1F). The calculated density (29) for the 1 D ice is1.038 g∕cm3, significantly higher than that of 0.9167 g∕cm3 forthe bulk ice at 0 °C (7).

When the guest water was fully removed, the nanochannel (1)retained its structure (Fig. S2), allowing measurement of the di-electric constants for (1) and water-free (1) (E ¼ 1.0 V, 10 kHz)from 45 to 298 K. As shown in Fig. 2, the real part of dielectricconstant (ϵ0) for (1) along the a axis exhibits a large temperaturedependence. It is nearly constant (ϵ0 ≈ 40) below 100 K but thenrises to a broad peak with a maximum ϵ0 ≈ 260 at around 170 K.As the temperature was further raised to 298 K, a sharp dielectricpeak appeared with the maximum ϵ0 ≈ 381 at around 277 K.Because the ϵ0 for (1) along the b and c axes and for water-free(1) along the a axis (ϵ0 ≈ 30) is almost temperature-independentover the entire range, we can conclude the temperature-depen-dent ϵ0 for (1) along the a axis corresponds to the water wire. Theϵ0 for the water wire is dramatically different from that of bulkwater [ϵ0ðH2OÞ ≈ 90 at 273 K] and exhibits high anisotropy com-pared to bulk water (30)

To establish the basis for the temperature-dependent beha-vior of the water wire, the dielectric constants for (1) and itsdeuterated form ½CuI2CuIIðCDTAÞð4;40-bpyÞ2�n • 6nðD2OÞ (1a)(Dataset S2) along the a axis were measured at various frequen-cies from 1 kHz to 10 MHz. As the frequency was raised through-out this range (Fig. 3A), the broad dielectric peak for (1) shifts tohigh temperature and its ϵ0 decreases from 293 to 82. At firstglance, the temperature- and frequency-dependent ϵ0 for (1)exhibits the characteristic behavior of a relaxor ferroelectric (thetypical strong dispersion effects often being ascribed to thefreezing-in of ferroelectric clusters; refs. 31 and 32). The broaddielectric peak, however, does not represent a phase transition,but a dielectric relaxation phenomena, because the broad peak ofϵ00 for (1) [especially that for (1a)] decreases rather than increaseswith the increase of frequency (33).

Consistent with this observation is that the temperature-dependent dielectric hysteresis revealed that no hysteretic polar-ization-electric field (P-E) loop was observed below 200 K whenmeasured at various ac electric fields of triangular waveformapplied along a crystallographic axis of the single crystal (Fig. 4).The ϵ0 for (1) decreases from 1,576 to 187 at 277 K (Fig. 3A),

significantly larger and sharper than the decrease in the broaddielectric peak. The sharp peak of 277 K clearly indicates adielectric phase transition (33). The significant spontaneouspolarization was observed only below 277 K and the satisfactoryP-E loop was obtained at approximately 250 K. The saturationpolarization (Psa) of ca. 2.5 μC·cm−2 and a residual polarization(Pr) of ca. 1.75 μC·cm−2 are obtained with a coercive field (Ec) ofca. 0.83 kV·cm−1.

Even though (1) exhibits a typical ferroelectric feature alongthe a axis, no P-E hysteresis loops are displayed along the b and caxes (Fig. S3). The strong anisotropy further affirms that the fer-roelectricity of (1) is attributed to the water wire. These results,together with the calculated density (29) for the water wire or1D ice [250 K(1b), 0.9946 g∕cm3; 260 K(1c), 1.0108 g∕cm3;278 K(1d), 1.0328 g∕cm3; and 298 K(1e), 1.0053 g∕cm3] (seeDatasets S3, S4, S5, and S6) obtained from crystal structuresat different temperatures, show that the dielectric phase transi-tion at 277 K is attributed to the transition between a 1D ferro-electric ice and a 1D liquid.

The temperature-dependent dielectric constants of (1a) werealso examined along the a axis at various frequencies in the rangeof 1–10 MHz (Fig. 3). Similar to that for (1), the broad dielec-tric peak of for (1a) also shifts to high temperature and its ϵ0decreases with increasing frequency, whereas the broad peakof ϵ0 for (1a) decreases with the increase of frequency. However,the maximum ϵ0 of the (1a) for the sharp peak appears at approxi-mately 285 K, which is approximately 8 K higher than that for (1),showing the “deuteration effect” that influences the phase transi-tion temperature of the water wire.

Given that (1) crystallizes with a the centrosymmetric spacegroup Fddd, what causes the ferroelectricity of the water wire?To address this question, the structure of water wire was analyzedin greater detail. Even though the positions of oxygen atoms inthe crystal structure can be fully determined from the X-raydiffraction (XRD) experiment, the exact positions of hydrogenatoms cannot. As mentioned above, however, O1w and O2w eachhydrogen-bond with O1 and O4 in the host nanochannel, therebyhindering the tendency for disordered hydrogen atoms. This hin-drance limits the orientation of remaining hydrogen atoms ofO1w and O2w, so they can only rotate along the axis of theO1⋯O1w and O4⋯O2w, respectively, to maximize hydrogen-bonding interactions in the water wire.

Ab initio molecular dynamics (AIMD) simulation (seeMaterials and Methods) was carried out to study the positionsof the water molecules within the nanochannel. We selectedone channel (containing 256 atoms, including 12 H2Omolecules)as the model system for the AIMD simulation. All the danglingbonds on the outer surface of the nanochannel were terminatedby hydrogen atoms. All the heavy atoms (except hydrogen) on thewall were fixed to the XRD measured positions, and they werenot relaxed during the geometry optimization and AIMD simula-tion. The initial orientations of hydrogen atoms were randomlyassigned. During the 12.5-ps AIMD simulation (with temperaturefixed at 300 K), we observed little movement in oxygen atompositions but considerable rotation of water molecules, whichresults in the continual formation and breaking of intermolecu-lar hydrogen bonds (see Movie S1). The oxygen atom positionsnear the end of the simulation are in good agreement with theexperimental values, as can be seen when comparing Fig. 1F witha snapshot of AIMD at 10 ps (Fig. 5A). In addition, based onthe optimized structure of the 12 water molecules within thenanochannel, we also estimated their net dipole moment whichamounts to approximately 9 D (or 0.75 D per water molecule).For this estimation, we used the simple-point-charge (SPC) watermodel with which each water has a dipole moment of 2.27 D(see Materials and Methods).

To gain more insight into the positions of hydrogen atomsunder an external electric field and to evaluate the magnitude

Fig. 2. The dielectric constant (ϵ0) of (1) measured at various temperaturesalong the a (black), b (red), and c (green) axes, respectively. The dielectricconstant (ϵ0) of water-free (1) measured at various temperatures along thea axis (blue) under the same condition (E ¼ 1 V, f ¼ 10 kHz).

Zhao et al. PNAS ∣ March 1, 2011 ∣ vol. 108 ∣ no. 9 ∣ 3483

APP

LIED

PHYS

ICAL

SCIENCE

S

Dow

nloa

ded

by g

uest

on

May

25,

202

1

Page 4: Transition from one-dimensional water to ferroelectric ice ...bonded to one O3w and one O2w) form a cyclic and boat-like water octamer. Each crystallographically independent water

of polarization under the external field, classical moleculardynamic simulation was employed (see Materials and Methods).In the molecular dynamics (MD) simulation, all the oxygen atomswere fixed at the positions determined by the measured crystalstructure while the orientations of O-H bonds were allowed torelax. From a snapshot of 1D ice under the external electricfield in the a direction, one can see that one hydrogen atom ofthe O1w forms a hydrogen bond either with symmetry-relatedO1w (Fig. 5B) or the symmetry-related O3w (Fig. 5C). Likewise,one hydrogen atom of O2w forms a hydrogen bond either withO3w (Fig. 5B) or symmetry-related O2w (Fig. 5C). For the O3w,one hydrogen atom either forms a hydrogen bond with O2w orsymmetry-related O3w. Note that the snapshots in Fig. 5 B and Cshow states in which most O-H bonds point along the electricfield, and that the snapshots in Fig. 5 B and C have the oppositeelectric field directions. The average dipole moment per 12 watermolecules in a channel along the a axis (Pa) is −25.1 and 25.0 D,respectively, in two opposite fields. Clearly, the hydrogen-bond-ing interactions among water molecules in the water wire and thehost (1) is of critical importance to the inherent ferroelectricityof the 1D ice. The water-host hydrogen bonds do not breakand reform while the remaining hydrogen atoms of the water mo-lecules in the 1D ice rotate alternately under the opposite electricfield. As a result, the polarity of the 1D ice can be reversedby reversing the external electric field, as shown in Fig. 5 Band C. Consistent with this structural explanation is that themagnitude of ϵ0 (about 1,576) with f ¼ 1 kHz at 277 K was nearly40 times larger than at 50 K (Fig. 4). The dielectric phase transi-tion is due to molecular motion rather than fast electronicmotion (34).

Water is the most abundant liquid on Earth and it plays a keyrole in many biological and chemical processes and materialsciences (35). However, its role in many phenomena is not fullyunderstood despite a myriad of studies (36–38). Here, basedon the confinement within the 3D supramolecular architectureof ½CuI2CuIIðCDTAÞð4;40-bpyÞ2�n, we have produced a 1D iceðH2OÞ12n. Single-crystal analysis revealed that the 1D ice, com-prised of 12 water molecules in a unit cell, can be viewed as awater tetramer that acts as a proton acceptor for a proton-donat-ing boat-like water octamer. Investigation on the temperature-dependent dielectric constant of (1) indicates that this water

wire exhibits large dielectric anomalies at around 175 and277 K, dramatically different from that of bulk water and ice XI(6–10, 37), and acts as a single-phase 1D ferroelectric ice below277 K. Analysis of the ferroelectric mechanism for the water wireshows that both dynamic hydrogen-bonding interactions amongwater molecules in the water wire and static interactions betweenwater molecules and the host play key roles. Hence, our studypoints out a viable route to the synthesis of ferroelectric materialsby taking advantage of the diversity and tunability of water withinmetal-organic frameworks.

Materials and MethodsSample Preparation. For ½CuI

2CuIIðCDTAÞð4;40-bpyÞ2�n • 6nH2O (1), trans-1,2-

diaminocyclohexane-N,N,N′,N′-tetraacetic acid monohydrate (H4CDTA,

Fig. 3. Temperature dependence of the complex dielectric constant (real part ϵ0 and imaginary part ϵ00) along a axis at different frequencies for (A) (1)and (B) (1a).

Fig. 4. The dielectric hysteresis of (1) along the a axis at different tempera-tures. (Insets) The side and top view of the 1D ice.

3484 ∣ www.pnas.org/cgi/doi/10.1073/pnas.1010310108 Zhao et al.

Dow

nloa

ded

by g

uest

on

May

25,

202

1

Page 5: Transition from one-dimensional water to ferroelectric ice ...bonded to one O3w and one O2w) form a cyclic and boat-like water octamer. Each crystallographically independent water

0.364 g, 1.0 mmol), copper acetate (CuðCH3COOÞ2 • H2O, 0.199 g, 1.0 mmol),4,4′-bipyridine (4,4′-bpy, 0.156 g, 1.0 mmol), and 10 mL distilled water weremixed while stirring at room temperature. When the pH of the mixture wasadjusted to about 3 with 1.0 mol·L−1 nitric acid, the solution was put into a25-mL Teflon-lined Parr, heated to 150 °C for 3 d, and then cooled to roomtemperature at a rate of 3 °C h−1. Red stick crystals of (1) were collected in ayield of 30% based on Cu. Analysis calculated for C34H46Cu3N6O14 (1) isas follows: C, 42.83; H, 4.86; N, 8.81; found: C, 42.99; H, 4.61; N, 8.68. IR(KBr pellet, cm−1): 3,431 s (strong); 3,044 w (weak); 2,929 m (moderate);2,858 w; 1,602 s; 1,531 w; 1,483 m; 1,408 s; 1,255 w; 1,216 m; 1,103 w;993 w; 916 w; 891 w; 819 m; 800 m; 727 w; 633 w; 570 w; 476 w.

Crystal data for (1) are as follows: Red crystal, 0.36 × 0.32 × 0.30 mm3,orthorhombic, space group Fddd with a ¼ 11.8037ð5Þ Å, b ¼ 24.8742ð12Þ Å,c ¼ 51.423ð3Þ Å, V ¼ 15;098.1ð12Þ Å3, Mr ¼ 953.39, Z ¼ 16, ρcalcd ¼1.678 g cm−3, μ ¼ 1.752 mm−1, F000 ¼ 7;856, T ¼ 123ð2Þ K, 2θmax ¼ 52.00°,RðintÞ ¼ 0.0312. Of the 10,335 reflections collected, 3,713 are independent re-flections (Rint ¼ 0.0312) and 2,745 are observed reflections [I > 2σðIÞ]. On thebasis of all these data and 255 refined parameters, R1ðobservedÞ ¼ 0.0441,wR2ðobservedÞ ¼ 0.1261, R1ðall dataÞ ¼ 0.0538, and wR2ðall dataÞ ¼ 0.1326were obtained, goodness of fit refine GOFð Þ ¼ 1.052.

Crystal data for (1b) are as follows: Red crystal, 0.36 × 0.32 × 0.30 mm3,orthorhombic, space group Fddd with a ¼ 11.7658ð5Þ Å, b ¼ 25.2769ð16Þ Å,c ¼ 51.469ð3Þ Å, V ¼ 15;307.1ð14Þ Å3, Mr ¼ 953.39, Z ¼ 16, ρcalcd ¼1.655 g cm−3, μ ¼ 1.728 mm−1, F000 ¼ 7;856, T ¼ 250ð2Þ K, 2θmax ¼ 52.00°,RðintÞ ¼ 0.0341. Of the 10,676 reflections collected, 3,676 are independent re-flections (Rint ¼ 0.0341) and 2,612 are observed reflections [I > 2σðIÞ]. On thebasis of all these data and 255 refined parameters, R1ðobservedÞ ¼ 0.0409,wR2ðobservedÞ ¼ 0.1111, R1ðall dataÞ ¼ 0.0653, and wR2ðall dataÞ ¼ 0.1194were obtained, GOF ¼ 1.003

Crystal data for (1c) are as follows: Red crystal, 0.36 × 0.32 × 0.30 mm3,orthorhombic, space group Fddd with a ¼ 11.7630ð6Þ Å, b ¼ 25.1910ð18Þ Å,c ¼ 51.501ð3Þ Å, V ¼ 15;260.9ð16Þ Å3, Mr ¼ 953.39, Z ¼ 16, ρcalcd ¼

1.660 g cm−3, μ ¼ 1.733 mm−1, F000 ¼ 7;856, T ¼ 260ð2Þ K, 2θmax ¼ 52.00°,RðintÞ ¼ 0.0369. Of the 10,633 reflections collected, 3,755 are independent re-flections (Rint ¼ 0.0369) and 2,521 are observed reflections [I > 2σðIÞ]. On thebasis of all these data and 255 refined parameters, R1ðobservedÞ ¼ 0.036,wR2ðobservedÞ ¼ 0.1153, R1ðall dataÞ ¼ 0.0719, and wR2ðall dataÞ ¼ 0.1245were obtained, GOF ¼ 1.015.

Crystal data for (1d) are as follows: Red crystal, 0.36 × 0.32 × 0.30 mm3,orthorhombic, space group Fddd with a ¼ 11.7584ð6Þ Å, b ¼ 25.0390ð19Þ Å,c ¼ 51.481ð3Þ Å, V ¼ 15;157.1ð17Þ Å3, Mr ¼ 953.39, Z ¼ 16, ρcalcd ¼1.671 g cm−3, μ ¼ 1.745 mm−1, F000 ¼ 7;856, T ¼ 277ð2Þ K, 2θmax ¼ 52.00°,RðintÞ ¼ 0.0405. Of the 10,619 reflections collected, 3,730 are independent re-flections (Rint ¼ 0.0405) and 2,428 are observed reflections [I > 2σðIÞ]. On thebasis of all these data and 255 refined parameters, R1ðobservedÞ ¼ 0.0473,wR2ðobservedÞ ¼ 0.1191, R1ðall dataÞ ¼ 0.0818, and wR2ðall dataÞ ¼ 0.1315were obtained, GOF ¼ 1.040.

Crystal data for (1e) are the following: Red crystal, 0.36 × 0.32 × 0.30 mm3,orthorhombic, space group Fddd with a ¼ 11.7658ð4Þ Å, b ¼ 25.2002ð17Þ Å,c ¼ 51.440ð2Þ Å, V ¼ 15;252.0ð13Þ Å3, Mr ¼ 953.39, Z ¼ 16, ρcalcd ¼1.661 g cm−3, μ ¼ 1.734 mm−1, F000 ¼ 7;856, T ¼ 298ð2ÞK, 2θmax ¼ 52.00°,RðintÞ ¼ 0.0287. Of the 9,254 reflections collected, 3,757 are independent reflec-tions (Rint ¼ 0.0287) and 2,745 are observed reflections [I > 2σðIÞ]. On the basisof all these data and 255 refined parameters, R1ðobservedÞ ¼ 0.0423,wR2ðobservedÞ ¼ 0.1060, R1ðall dataÞ ¼ 0.0718, and wR2ðall dataÞ ¼ 0.1115were obtained, GOF ¼ 1.011.

½CuI2Cu

IIðCDTAÞð4;40-bpyÞ2�n • 6nD2O (1a) was prepared in a similar way asdescribed for (1), expect for using 10 mL deuterium oxide to replace water.Red crystals were obtained in a 25% yield based on Cu. Analysis calculated forC34H34D12Cu3N6O14 (1a) is as follows: C, 42.30; N, 8.70; found: C, 43.52; N,8.66. IR (KBr pellet): 3,428 s; 3,073 w; 3,043 w; 2,932 w; 2,864 w; 1,601 s;1,528 w; 1,483 m; 1,408 s; 1,215 m; 1,137 w; 1,063 w; 993 w; 954 w;935 w; 875 m; 818 m; 800 m; 724 w; 633 w; 570 w; 476 w. Crystal data for1a are red crystal, 0.30 × 0.15 × 0.08 mm3, orthorhombic, space group Fdddwith a ¼ 11.7434ð19Þ Å, b ¼ 24.892ð4Þ Å, c ¼ 50.997ð8Þ Å, V ¼ 14;907ð4Þ Å3,Mr ¼ 965.46, Z ¼ 16, ρcalcd ¼ 1.721 g cm−3, μ ¼ 1.774 mm−1, F000 ¼ 7;856,T ¼ 123ð2Þ K, 2θmax ¼ 52.00°, RðintÞ ¼ 0.0295. Of the 19,439 reflectionscollected, 3,654 are independent reflections (Rint ¼ 0.0295) and 3,286 areobserved reflections [I > 2σðIÞ]. On the basis of all these data and 258 refinedparameters, R1ðobservedÞ ¼ 0.0509, wR2ðobservedÞ ¼ 0.1331, R1ðall dataÞ ¼0.0557, and wR2ðall dataÞ ¼ 0.1296 were obtained, GOF ¼ 1.042.

Crystal Structure Determination. Data were collected with one single crystal onOxford Gemini S Ultra using Mo-Kα radiation (λ ¼ 0.71073) and equipped withOxford Instruments at 123 K for (1), 250 K for (1b), 260 K for (1c), 277 K for (1d),and 298K for (1e) (Datasets S1, S3, S4, S5, and S6). Data of (1a) were collected ona Bruker SMART Apex CCD diffractometer using graphite-monochromatizedMo-Kα radiation (λ ¼ 0.71073) at 123K (Dataset 2). Absorption correctionswereapplied using the multiscan program SADABS. The structures were solvedby direct methods (SHELXTL, version 5.10), and the non-H-atoms were refinedanisotropically by a full-matrix least-squares method on F2. CCDC 800743–800748 contain the supplementary crystallographic data for this paper as well.These data canbeobtained freeof charge from theCambridgeCrystallographicData Centre via www.ccdc.cam.ac.uk/data_request/cif.

Physical Measurements. Temperature-dependent dielectric constants weremeasured using the two-probe ac impedance method for frequencies from1 kHz to 10 MHz (Wayne Kerr 6500B). A single crystal was placed into acryogenic refrigeration system (Oxford Cyromech). The electrical contactswereprepared using gold paste (Tokuriki 8560) to attach 25-μm gold wires to thesingle crystal. The P-E curve was measured using Radiant Precision Premier II.

Ab Initio Molecular Dynamics Simulation. Ab initio (Born-Oppenheimer)molecular dynamics simulation was performed using the QUICKSTEP pro-gram implemented in the CP2K package (39, 40). The Becke-Lee-Yang-Parr(41, 42) exchange correlation functional was selected, and the empirical dis-persion correction (43, 44) to the density-functional theory (DFT) calculation(i.e., DFT-D calculation) was used to model the intermolecular dispersion. TheGoedecker-Teter-Hutter (GTH) (45) norm-conserved pseudopotentials wereutilized to account for the contribution of the core electrons. We also usedthe GTH-double-zeta-polarization Gaussian basis set along with a plane-wave basis set (with an energy cutoff of 280 Ryd) for the expansion of elec-tronic wavefunctions. The AIMD simulation was performed in a constant-vo-lume and constant-temperature ensemble with the temperature controlledat 300 K. The simulation cell contains 256 atoms, including 12 water mole-cules. The total AIMD simulation time was 12.5 ps (Movie S1).

Fig. 5. (A) A snapshot of water molecules (plotted in a ball and stick model)and vicinal atoms of nanochannel (plotted in a line model) at 10 ps from anAIMD simulation with temperature controlled at 300 K. Classical MD snap-shots of the 1D ice in an external electric field whose direction is (B) alongthe −a axis and (C) along the a axis, respectively. In B and C, the red and greenballs denote to O and H atoms, respectively, and the orange dashed linesdenote hydrogen bonds.

Zhao et al. PNAS ∣ March 1, 2011 ∣ vol. 108 ∣ no. 9 ∣ 3485

APP

LIED

PHYS

ICAL

SCIENCE

S

Dow

nloa

ded

by g

uest

on

May

25,

202

1

Page 6: Transition from one-dimensional water to ferroelectric ice ...bonded to one O3w and one O2w) form a cyclic and boat-like water octamer. Each crystallographically independent water

Classical Molecular Dynamics Simulation. The classical MD simulations wereperformed using the Discover program implemented in the Material Studio4.4 package (46). The consistent-valence force field was used with the atomiccharges derived from the charge equilibration (47) method for the supramo-lecular structure, and the SPC (48) model was used for water. The simulationcell contains 1,648 atoms, including 96 water molecules. The initial structurewas equilibrated at 290 K for 1 ns. After the equilibration, two additional MDsimulations were performed, one in the external electric field (5.14 V∕nm)along the 1D water wire, and another in opposite direction.

ACKNOWLEDGMENTS. X.C.Z. is grateful to Prof. Mark Griep for his criticalreading of the manuscript. Xiamen University group is supported by theNational Natural Science Foundation of China (Grants 20825103, 20721001,and 90922031) and the 973 project (Grant 2007CB815304) from the Ministryof Science and Technology of China. The University of Nebraska-Lincolngroup is supported by the National Science Foundation (Grants CHE-0701540 and CBET-1036171), Army Research Office (Grant W911NF1020-099), Nebraska Research Initiative, and University of Nebraska’s HollandComputing Center.

1. Hummer G, Rasaiah JC, Noworyta JP (2001) Water conduction through the hydropho-bic channel of a carbon nanotube. Nature 414:188–190.

2. Sansom MSP, Biggin PC (2001) Water at the nanoscale. Nature 414:156–159.3. Levinger NE (2002) Water in confinement. Science 298:1722–1723.4. Cheruzel LE, et al. (2003) Structures and solid-state dynamics of one-dimensional

water chains stabilized by imidazole channels. Angew Chem Int Ed 42:5452–5455.5. Kolesnikov AI, et al. (2004) Anomalously soft dynamics of water in a nanotube:

A revelation of nanoscale confinement. Phys Rev Lett 93:035503-1–035503-4.6. Bramwell ST (1999) Ferroelectric ice. Nature 397:212–213.7. Petrenko VF, Whiteworth RW (1999) Physics of Ice (Oxford Univ Press, New York).8. Fukazawa H, Hoshikawa A, Ishii Y, Chakoumakos BC, Fernandez-Baca JA (2006)

Existence of ferroelectric ice in the universe. Astrophys J 652:L57–L60.9. Fukazawa H, Hoshikawa A, Chakoumakos BC, Fernandez-Baca JA (2009) Existence of

ferroelectric ice on planets—a neutron diffraction study. Nucl Instrum Methods PhysRes, Sect A 600:279–281.

10. Iedema MJ, et al. (1998) Ferroelectricity in water ice. J Phys Chem B 102:9203–9214.11. Su X, Lianos L, Shen R, Somorjai GA (1998) Surface-induced ferroelectric ice on Pt(111).

Phys Rev Lett 80:1533–1536.12. Jackson SM, Whitworth RM (1997) Thermally-stimulated depolarization studies of the

ice XI-Ice Ih phase transition. J Phys Chem B 101:6177–6179.13. Luo CF, Fa W, Zhou J, Dong JM, Zeng XC (2008) Ferroelectric ordering in ice nanotubes

confined in carbon nanotubes. Nano Lett 8:2607–2612.14. Mikami F, Matsuda K, Kataura H, Maniwa Y (2009) Dielectric properties of water inside

single-walled carbon nanotubes. ACS Nano 3:1279–1287.15. Lines ME, Glass AM (1977) Principles and Applications of Ferroelectrics and Related

Materials (Oxford UnivPress, New York).16. LeBard DN, Matyushov DV (2010) Ferroelectric hydration shells around proteins:

Electrostatics of the protein/water interface. J Phys Chem B 114:9246–9258.17. Cui HB, et al. (2008) A porous coordination-polymer crystal containing one-

dimensional water wires exhibits guest-induced lattice distortion and a dielectricanomaly. Angew Chem Int Ed 47:3376–3380.

18. Bai JE, Wang J, Zeng XC (2006) Multiwalled ice helixes and ice nanotubes. Proc NatlAcad Sci USA 103:19664–19667.

19. Bishop Cl, et al. (2009) On thin ice: Surface order and disorder during pre-melting.Faraday Discuss 141:277–291.

20. Maniwa Y, et al. (2002) Phase transition in confined water inside carbon nanotubes.J Phys Soc Jpn 71:2863–2866.

21. Ghosh S, Ramanathan KV, Sood AK (2004) Water at nanoscale confined insingle-walled carbon nanotubes studied by NMR. Europhys Lett 65:678–684.

22. Maniwa Y, et al. (2005) Ordered water inside carbon nanotubes: Formation ofpentagonal to octagonal ice-nanotubes. Chem Phys Lett 401:534–538.

23. Byl O, et al. (2006) Unusual hydrogen bonding in water-filled carbon nanotubes. J AmChem Soc 128:12090–12097.

24. Koga K, Gao GT, Tanaka H, Zeng XC (2001) Formation of ordered ice nanotubes insidecarbon nanotubes. Nature 412:802–805.

25. Febles M, et al. (2006) Distinct dynamic behaviors of water molecules in hydratedpores. J Am Chem Soc 128:10008–10009.

26. Cheruzel LE, et al. (2003) Structures and solid-state dynamics of one-dimensionalwater wires stabilized by imidazole channels. Angew Chem Int Ed 42:5452–5455.

27. Janiak C, Scharamann TG (2002) Two-dimensional water and ice layers: Neutrondiffraction studies at 278, 263 and 20 K. J Am Chem Soc 124:14010–14011.

28. Long LS, Wu YR, Huang RB, Zheng LS (2004) A well-resolved uudd cyclic water tetra-mer in crystal host of [Cu(adipate)(4,4-bipyridine)] ðH2OÞ2. Inorg Chem 43:3798–3800.

29. Spek AL (2003) Single-crystal structure validation with the program PLATON. J ApplCryst 36:7–13.

30. Murrell JN, Jenkins AD (1994) Properties of Liquids and Solutions (Wiley, Chichester,England), 2nd Ed.

31. Samara GA (2003) The relaxation properties of compositionally disordered ABO3

perovskites. J Phys Condens Matter 15:R367–R411.32. Cross LE (1987) Relaxor ferroelectrics. Ferroelectrics 76:241–267.33. Horiuchi S, Kumai R, Tokunaga Y, Tokura Y (2008) Proton dynamics and room-

temperature ferroelectricity in anilate salts witha proton sponge. J Am Chem Soc130:13382–13391.

34. Kao KC (2004) Dielectric Phenomena in Solids (Elsevier, San Diego).35. Pratt LR (2002) Introduction: Water. Chem Rev 102:2625–2626.36. Zubavicus Y, Grunze M (2004) New sights into the structure of water with ultrafast

probes. Science 304:974–976.37. Singer SJ, et al. (2005) Hydrogen-bond topology and the ice VII/VIII and ice Ih/XI pro-

ton-ordering phase transitions. Phys Rev Lett 94:135701.1–135701.4.38. Run CY, Lobastov VA, Vigliotti F, Chen S, Zewail AH (2004) Ultrafast electron

crystallography of interfacial water. Science 304:80–84.39. VandeVondele J, et al. (2005) Quickstep: Fast and accurate density functional

calculations using amixed Gaussian and plane waves approach. Comput Phys Commun167:103–128.

40. Lippert G, Hutter J, Parrinello M (1997) A hybrid Gaussian and plane wave densityfunctional scheme. Mol Phys 92:477–487.

41. Becke AD (1988) Density-functional exchange-energy approximation with correctasymptotic behavior. Phys Rev A 38:3098–3010.

42. Lee C, Yang W, Parr RG (1988) Development of the Colle-Salvetti correlation-energyformula into a functional of the electron density. Phys Rev B 37:785–789.

43. Grimme S (2004) Accurate description of van der Waals complexes by densityfunctional theory including empirical corrections. J Comput Chem 25:1463–1473.

44. Grimme S (2006) Semiempirical GGA-type density functional constructed with along-range dispersion correction. J Comput Chem 27:1787–1799.

45. Hartwigsen C, Goedecker S, Hutter J (1998) Relativistic separable dual-space Gaussianpseudopotentials from H to Rn. Phys Rev B 58:3641–3662.

46. Materials Studio (Accelrys, San Diego) Version 4.4.47. Dauber-Osguthorpe P, et al. (1988) Structure and energetics of ligand binding to

proteins: E. coli dihydrofolate reductase-trimethoprim, a drug receptor system.Proteins: Struct Funct Genet 4:31–47.

48. Rappé AK, Goddard WA, III (1991) Charge equilibration for molecular dynamicssimulations. J Phys Chem 95:3358–3363.

3486 ∣ www.pnas.org/cgi/doi/10.1073/pnas.1010310108 Zhao et al.

Dow

nloa

ded

by g

uest

on

May

25,

202

1