You are on page 1of 151

THE ANOMALOUS PROPERTIES OF LIQUID WATER

EXPLAINED BY A MIXTURE MODEL

by

MARY S. VEDAMUTHU, B.S., M.S.

A DISSERTATION

IN

CHEMISTRY

Submitted to the Graduate Faculty


of Texas Tech University in
Partial Fulfillment of
the Requirements for
the Degree of
DOCTOR OF PHILOSOPHY

May, 1996

it>lWVj*
-f\C s^^
PKX'

1996, Mary Vedamuthu


ACKNOWLEDGEMENTS

I am deeply indebted to Professor G. Wilse Robinson for his immense support and

guidance throughout this work. His enthusiasm for science was both contagious and an

inspiration to me. I thank him for giving me the opportunity to learn and to expand my

scientific experience.

I must express my most sincere gratitude to Dr. Surjit Singh for his undying

enthusiasm and encouragement and for bestowing upon me a sense of the adventure for

science. His mathematical expertise has been invaluable to the theoretical aspects of this

work.

I would like to thank Professor Edward L. Quitevis, Professor Richard L. Redington

and Professor Richard E. Wilde for their helpful discussions and guidance throughout the

program. I would also like to express my thanks to Professor Robert D. Walkup and

Professor Edward L. Quitevis, who, as graduate advisors gave me many valuable suggestions

and advice, which made my stay here successful and enjoyable.

I would like to acknowledge the help and cooperation I have received from my

colleagues. Dr. Ningyi Luo, Mr. Chul Hee Cho and Mr. Jacob Urquidi. I also thank the

Department of Chemistry and Biochemistry, the R.A. Welch Foundation and the G. Wilse

Robinson, Jr. Trust for their financial support.

I am greatly indebted to my husband, parents, brother, sister andfriendsfor their

continued support and encouragement without which this work could not have been

completed.

11

m^
3>
- ' - ' ' '

TABLE OF CONTENTS

ACKNOWLEDGMENTS 11

ABSTRACT . VI

LIST OF TABLES Vlll

LIST OF FIGURES

CHAPTER
1
I. INTRODUCTION .
3
1.1. Significance of Ice Polymorphism
5
1.2. Structural Aspects of the Ice Phases
8
1.2.1. Low-Pressure Ice Phases 10
1.2.2. High-Pressure Ice Phases 16
1.3. Structural Models for Liquid Water 16
23
1.3.1. Survey of Early Work
1.3.2. More Recent Models 26
1.4. Interfacial Water
1.5. Present Work 29

n. THE DENSITY ANOMALY OF LIQUID H2O EXPLAINED


BY THE MIXTURE-MODEL . . . . 32

2.1. Qualitative Description of the Density Anomaly 34

2.2. Experimental Density Data 35

2.3. Density Analysis . . . . 38

2.4. Density Fits . . . . . 45

111

Mm
2.4.1. Exact Algebraic Fits . . . . 46

2.4.2. Least-Squares Fits . . . . 52

2.5. Other Fits . . . . 62

2.6. Discussion . . . . 68

III. ACCURATE MIXTURE-MODEL DENSITIES FOR D2O 70

3.1. Experimental Density Data . . . . 70

3.2. Density Fits . . . . . . 73

3.2.1. Preliminary Fits . . . . . 74


3.2.2. Exact Algebraic Fit . . . . 75
3.2.3. Least-Squares Fit . . . . 75

3.3. Discussion . . . . . . 76

IV. A SIMPLE RELATIONSHIP BETWEEN THE DENSITIES


AND OTHER PROPERTIES OF ISOTOPIC WATER 83

4.1. Comparison of the H2O and D2O Densities Using the

Structural Relationship . . . . . 83

4.2. ScalingProcedure Applied to the Density of T2O 88

4.3. Application to Other Properties . . . . 91

V. THE ISOTHERMAL COMPRESSIBILITY MINIMUM NEAR

50 C AND OTHER ANOMALOUS PROPERTIES 93

5.1. Qualitative Description of the Isothermal Compressibility 93

5.2. Quantitative Description of the Isothermal Compressibility 95

5.2.1. Experimental data . . . . 96


5.2.2. The Fitting Procedure . . . . 100
5.2.3. Discussion . . . . . 105

IV
5.3. A Note about the Heat Capacity, Dynamic Properties
and Viscosity 107

VI. INDEPENDENT SUPPORT FOR THE MIXTURE MODEL 109

6.1. Radial Distribution Function 109

6.2. Kamb'sWork 110

6.3. Internal Vibrational Frequencies 111

6.4. Singular Behavior near To 113

6.5. Structural Relationships . 113

6.6. Isochoric Temperature Differential of the


X-Ray Structure Factor . 114

6.7. Computational Studies . 116

VII. LASER STUDIES OF THE PROPERTIES OF LIQUID


WATER PERTURBED BY SURFACES 119

7.1. Experimental Section 120

7.2. Analysis of Water Confined in Small Volumes 122

7.2.1. Interfacial Water in Thin Films . 122


7.2.2. Water Confined Between Quartz Plates 125

7.3. Conclusions 126

VIII. CONCLUSIONS 127

REFERENCES 130

APPENDIX:
LEAST SQUARES FITTING PROCEDURE 140
ABSTRACT

A semiempirical theory, which is a modernization of the 'mixture model,' attempting

quantitatively to correlate thermodynamic and dynamic effects of bulk and interfacial liquid

water with various properties of ice polymorphs is proposed here. The basic concept rests

on the disappearance, on the average, with increasing temperature or pressure, of open

intermolecular tetrahedral bonding (Type-I) having a density similar to that of ordinary ice,

in favor of compacdy bonded regions (Type-II) with a density near that of the dense ice

polymorphs particularly, ice II, III, and V.

The mixture model is employed to explain quantitatively the origins of the

'anomalous' properties of liquid water - density maximum, isotope effects, thermal minimum

in the isothermal compressibility curve. Strong support for this model can be found from

an analysis of the accurate experimental density data of liquid H2O and D2O from the

supercooled regime to about +70 C. Published density data can be fit to this model with

six- to seven-decimal-point accuracy, in the case of liquid H2O and to the reported five-

decimal-point precision, in the case of liquid DjO. The output parameters from the fits

indicate the presence of capacious intermolecular bonding with a density extremely close to

that of ordinary ice-Ih, intermixed with compactly bonded regions having a density near that

of the common dense forms of ice. A quantitative assessment of the temperature dependence

of the isothermal compressibility of liquid water at atmospheric pressure was carried out.

The 'anomalous' minimum in this quantity near 50 C is shown to emerge naturally.

Independent support for this model has been provided by the differential x-ray scattering

VI

^,*SSSS>aw*jj>.i.
experiments of Bosio et al. Their resuhs clearly indicate that a dynamic, temperature

dependent mixture of ice-I-, -II-, -III-, and -V-type bonding is present in the liquid in the

manner expected for the model described in this work. Based on eariier x-ray scattering

studies, Kamb reached a similar conclusion about these mixed bonding forms in liquid water.

Recently, computational studies conducted by Cho, Singh and Robinson in our laboratory

have indicated that the density anomaly of liquid water can be explained by utilizing this

mixture model concept.

Ultrafast laser methods were used to analyze the properties of liquid water confined

in small volumes. This study shows that interfacial water appears to be more structured and

orientationally stiffer than bulk water.

Vll
LIST OF TABLES

1.1. Structural Features ofthe Ice Phases . . . 6

1.2. Bonding Characteristics of the Ice Phases 7

2.1. Liquid Water Densities (g cm'^) at 1 atm from


Kell's Equation Compared with Experimental
Supercooled Densities 39

2.2. A Few Forms off, (T) Tested for the Density Analysis . 42

2.3. Liquid H2O Densities (g cm'^) at 1 atmfromKell's


EquationCompared with Fit (2.1) Densities 49

2.4. Fitting Parameters from Fits (2.1) and (2.2) 51

2.5. Fitting Parameters from Fit (2.3) . . . 54

2.6. Fitting Parameters from Fits (2.4) and (2.5) 57

2.7. Fitting ParametersfromFit (2.6) . . . 60

2.8. Fitting ParametersfromFit (2.7) . . . 61

2.9. Recalculated Densities for the Least-Squares Fits (2.4),


(2.6) and (2.7) Compared with thosefromKell Data . 63

2.10. Temperature-Dependent Fractions of Ice-I-Type

Bonding Fit (2.1) Compared with those from Fit (2.6) 64

2.11. FittingParameters from Fits (2.8) and (2.9) 66

2.12. Fitting Parameters from Fits (2.10) and (2.21) . 67


3.1. Liquid D2O Densities (g cm"^) at 1 atmfromKell's
Equation Compared with Experimental H-S
Supercooled Densities . . . . . 72

3.2. Fitting Parameters from Fits (3.1) and (3.2) 77

3.3. Liquid D2O Densities (g cm"^) at 1 atmfromKell's Equation


Compared with Fit (3.1) and Fit (3.2) Densities 78

Vlll
3.4. Recalculated Densities for Fits (3.1) and (3.2)
Compared with those from Kell . . . . 79

3.5. Temperature-Dependent Fractions of Ice-I-Type Bonding

from Fit (2.1) Compared with those from Fit (3.1) 80

4.1. Unsealed Densities of DjO . . . . 86

4.2. Scaled Densities of D2O . . . . 87

4.3. Unsealed Densities of TjO . . . . 89

4.4. Scaled Densities of T.O 90


5.1. Observed and Calculated Isothermal Compressibility
Factors as a Function of Temperature 98

5.2. The Volume Quantities in Eq.5.2 as a


Function of Temperature 103

5.3. Volume Parametersfromthe Density Analysis and Pressure


Parametersfromthe Least-Squares Fit (5.1) 104

5.4. The Three Contributing Terms and dfjd? in Eq.5.2 as a


Function of Temperature . . . . 106

IX

,1v^4iVL*5-^VV.,"^" '.i^
LIST OF FIGURES

1.1. Crystal Structure of Ice-Ih 9

1.2. Crystal Structure of Ice-II 13

7.1. 1-Naphthol Decay Lifetime in Water Confined


in an Ultra-thin Film 124

ES^'V.ir?'"':^?*.,'^' 5
CHAPTER I

INTRODUCTION

Water covers nearly three-fourths of the earth's surface. It is present in the

atmosphere and the earth's crust and composes a large part of all living plant and animal

matter. Nutrients are transported into the roots of plants as solutions in water. Biological

reactions occur in solution in the water in cells. Many industrial reactions are run in water

and without water many chemical reactions would not take place. Water has its own

chemical behavior, and it has been used as a standard for many physical constants and units.

However, in what manner water plays a role in the operation of a chemical or biological

system is still a matter of debate.

Water is often called the universal solvent [1]. A large variety of chemical reactions

take place in water or in aqueous mixed solvents. These aqueous media often display

amazing solvent effects on the rate and thermodynamic activation parameters of many

organic reactions [2]. Many enzymes catalyze biochemical reactions through water

molecules bound to their active sites. The explicit inclusion of interfacial water is absolutely

essential in the unveiling of many problems in biology and biochemistry [3-8]. The

interaction of water with clays, surfactants, and metal surfaces is important in

technologically important concerns such as oil recovery, mining, catalysis, corrosion

inhibition, etc. [9]. Aftindamentalknowledge of water is essential to understand fully the

different functions of this unique liquid.

gy.^i^y.'i,^iV,ii -^ - j ^
Water displays a striking set of physical and chemical properties [10-16], and some

of them are apparently unique. Some ofthe more important attributes are given below:

1. Contraction of ice on melting at 0 C and atmospheric pressure, a property relatively rare

among all substances.

2. Density maximum in the liquid at 3.984 C [17], no other liquid is known with a

corresponding density maximum above its normal melting point. By increasing the

external pressure, the maximum of water density shifts downward reaching 0 "C at 190

bars.

3. Isothermal compressibility decreases with increasing temperature from its melting point

and reaches a minimum at 46.5 C [18]. However, the anomalous phenomenon disappears

at pressures above 3 kbars when the compressibility increases with temperature as in

normal liquids.

4. Negative pressure coefficient of viscosity for temperatures below about 20 ''C at not too

high a pressure [19], which means that, unlike other liquids, increasing the pressure lowers

the viscosity of liquid water.

5. Molar heat capacity of water higher than the value expected from standard considerations

of the contributions from various degrees of freedom.

6. Characteristic temperatures such as melting, boiling, and critical temperatures

anomalously high, unlike other substances having comparable molecular weights.

7. Large dielectric constant and a high surface tension.

KJ^iMlPPIUff'aia
8. The muhiplicity of ice crystal structures exhibited by water [20] suggests that water

molecule interactions must be rather complicated as it is highly unlikely that spherical

molecules, interacting through simple potentials, could exhibit as many crystal structures.

Although life on this planet has been conditioned by the abnormal properties of

water, many of these properties have not had a well-accepted explanation. The existence of

these unusual properties adds extra zest to the task of developing a viable theory of water,

which uhimately is charged with connecting these properties to molecular structure and

interactions. Therefore, probing into the structural properties of liquid water has a

ftindamental utility in many areas of pure science, engineering and industry.

1.1. Significance of Ice Polvmorphism in


Relation to the Structure of Water

One ofthe best examples of 19th century musings about the properties of liquid

water was recorded by W. C. Rontgen [21]. With the development of notions of chemical

equilibrium in the nineteenth century, it came to be realized that the maximum density of

water can be explained by an equilibrium between two types of differently constructed

molecules. Rontgen called the first type 'ice molecules' for they possessed certain

characteristics of ice. The transformation from molecules offirsttype to those ofthe second

type on heating resulted in a decrease in the volume. Rontgen [21] sought to extend the idea

to explain the decrease of compressibility from 0 to 50 "C, the increase of thermal expansion

with increasing pressure in the same temperature range, and the decrease of viscosity with

increasing pressure. He argued qualitatively that the concentration of dissolved ice

molecules must be less at high temperatures and high pressures, which explains the
phenomena. Although Rontgen in his 1892 paper talks about a 'second type of molecule'

which was bulkier than the first type, he cleariy knew very little about this. The

crystallography ofthe polymorphs of ice originated with Bridgman [22] about twenty years

later. In 1900, Sutheriand put forward a theory [23], representing water as a mixture of

monohydrol (monomeric, steamlike), dihydrol (dimeric, intrinsically liquid) and trihydrol

(trimeric, ice-like) molecules. This was a particularization ofthe Rontgen mixture concept,

but was attempted too soon, so that, although it dominated the scene for more than a decade,

it ended up being completely discredited by fundamental developments in valence and

structure theories.

Since the time of Rontgen, the relation ofthe structure of water to the structure of ice

has been pointed out often. In 1926, Tammann [24] suggested that there should be as many

different 'kinds' of water as there are different phases of ice. Following this, in 1933, Bemal

and Fowler [25] were able to recognize the important role of hydrogen bonding in the

structure and properties of water. They explained the increase in density on mehing of ice

to water by proposing an analogy between water and quartz, the density increase on going

from ice -I to water being the same asfromtridymite to quartz. This proposal introduced

the idea of hydrogen-bond bending in water, since the Si-O-Si linkage in quartz is bent

through an angle of about 35 degrees. Bemal and Fowler pointed out that, if this idea were

correct, there should occur among the dense polymorphs of ice a structure analogous to that

of quartz. Later, in 1938, Morgan and Warren compared the radial distribution function

(r.d.f) of water at +4 "C with r.d.f's for ice polymorphs and hypothetical ice structures [26],

;^.?^T^v4.?';^^
They showed that a structure analogous to quartz does not occur among the dense forms of

ice and provided evidence against the proposed quartz-like structure for liquid water.

It is now widely recognized that a basic structural relation exists between the two

types of condensed phases, liquid and solid, and that the structural change in melting is

primarily a loss of crystalline long-range order, accompanied by a great increase in

molecular or atomic mobility [27]. It is also known that for substances exhibiting high-

pressure polymorphism, the liquid phase is able to anticipate structurally the features of

dense polymorphs and the labile structure of the liquid incorporates features of dense

polymorphs. When the dense contribution from a high-pressure polymorph is sufficiently

abundant, the change of volume on melting is negative over a range of pressures below the

solid-solid phase transition. The density increase on mehing of ordinary ice is directly

traceable to the fact that, above 2 kbar ice shows a high-pressure polymorphism. The

structural change responsible for the increase in density is basically the same as what occurs

in the transition from ice -Ih to the dense ice phases. Thus, ice polymorphism has an

important bearing on liquid water structure.

1.2.Structural Aspects ofthe Ice Phases

The distinct ice phases of H2O are nine in number which are listed in Tables 1.1 and

1.2 with characterizing physical and structural information. Ordinary ice -Ih is the only

phase of ice stable at low pressures. Ice -Ic, sometimes called cubic ice, has no region of

actual stability, being formed only as a metastable phase and at about -90 "C it inverts to ice

-Ih.

waSB^'ii! ^ "'> i

H9PI![B!HnHIHPM|P
Table 1.1 Structural Features ofthe Ice Phases

Ice Phase Crystal System Z" Density'' Pressure Temperature


gcm"^ kbar "C
I Hexagonal 4 0.92 0.0 0

Ic Cubic 8 0.93 0.0 -87

II Rhombohedral 12 1.18 2.1 -35

III Tetragonal 12 1.15 2.1 -30

V Monoclinic 28 1.24 3.4 -20

VI Tetragonal 10 1.34 8.0 15

VII Cubic 2 1.65 25.0 25

VIII Cubic 2 1.66 25.0 -50

IX Tetragonal - 1.16 2.0 -163

Sources of data are given in Refs. 29 and 30.

Number of molecules in the unit cell.

Density is at the pressure and temperature listed.

f\-n,':xj^^mimtimm^^'
Table 1.2 Bonding Characteristics ofthe Ice Phases

Ice Phase Bond' Bond Length'' Next-nearest Bond Bending*^


Frameworks A Neighbor Distance deg^xlO^
A
Ih 2.75 4.5 0

Ic 2.75 4.5 0

II 2.80 3.24 0.29

III 2.78 3.47 0.27

V 2.80 3.28, 3.46 0.34

VI 2 2.81 3.51 0.53

VII 2 2.96 -2.95 0

VIII 2 2.96 -2.95 0.01

IX 1 2.78 3.46 0.27

Sources of data are given in Refs. 29 and 30.

"Number of independent hydrogen-bond frameworks.

''Mean 00 distance for hydrogen bonds.

* Mean square deviation of 0 - 0 - 0 bond anglesfromtetrahedral.


The high pressure polymorphism of ice starts at about 2 kbar. All the high pressure phases

are stable under the conditions indicated, except ice -IV, whose existence is at best

metastable and transitory, its metastable equilibrium with ice -VI and with its liquid has been

observed [28].

1.2.1. Low-Pressure Ice Phases

The well-known ice -Ih structure, in which each molecule is hydrogen bonded to four

others in nearly perfect tetrahedral coordination, serves as a basis for the concept of water

as a tetrahedral molecule. In this concept, the hydrogen bond is formed primarily by

electrostatic attraction between a proton of one molecule and an unshared electron pair of

another, the tetrahedral bond directionality being a consequence of an approach to sp^ orbital

hybridization in the electronic structure ofthe water molecule. In the structure of ice -Ic, the

water molecule coordination is constrained to be perfectly tetrahedral by the cubic symmetry

ofthe crystal.

The structure ofthe low-pressure form, ice-Ih, is shown in Figure 1.1. Each water

molecule is hydrogen-bonded to four near neighbors with an 0-H--O distance of 2.75 A.

The H-O-H angle ofthe isolated water molecule, 104.6, matches fairiy closely the ideal

tetrahedral coordination angle of 109.5. The protons lie asymmetrically in each hydrogen

bond, about 1.01 Afromone oxygen atom and 1.74 Afromthe other, and each oxygen atom

has only two nearby protons, so each oxygen site corresponds to an intact water molecule.

In ice -Ih and -Ic, the next-nearest neighbors lie at a distance of 4.50 A. The next-nearest

8
' '" - f^c^l'. - ^

Figure 1.1. Crystal Structure of Ice-Ih


neighbor coordinations have significance for liquid structure models as do the nearest

neighbor coordinations.

1.2.2. High-Pressure Ice Phases

Eariier studies [29,30] ofthe dense forms of ice suggested that the structure ofthe

water molecule had been broken down and replaced with an arrangement of H' and 0'' ions.

Later work [31,32] shows that all the forms of ice are hydrogen bonded, although the

strength ofthe hydrogen bonding generally decreases with increasing density. In all ofthe

ice phases, the water molecules maintain their individuality almost unchanged from their

configuration in water vapor, and each molecule forms four hydrogen bonds, making a bond

framework that traverses the entire crystal. The early idea that the different forms of ice

correspond to different water polymers [24] is not really valid, in the sense that different

discrete, tightly bonded groups of molecules cannot be singled out in the actual ice

structures.

The bonding coordination in the dense ice phases is tetrahedral in the sense that each

molecule forms four hydrogen bonds of approximately equal strength. But, by contrast with

ices -Ih and -Ic, the arrangement ofthe bonded neighbors is more or less distorted from the

ideal tetrahedral geometry. This distortion represents a bending of the hydrogen bonds,

because of the unchanging geometry of the individual water molecules. However, the

hydrogen-bonding interactions are not strong enough to cause any significant rehybridization

that would allow a change in the H-O-H angle or in the spatial distribution ofthe unshared

10

US^M^Oii.'-,
electron pairs. A simple measure of bond bending is the mean square deviation of the

0 - 0 - 0 bond angles from 109.5".

The extent of bond bending increases progressively with increased density (except

for ices -VH and -VIII) (see Table 1.2). The distortions from ideal tetrahedral coordination

around each water molecule make possible increased densities by allowing the

accommodation of next-nearest neighbors in the distance range 3 to 4 A. This distance is

sufficient for the next-nearest neighbors to interact significantly with the central molecule

by dispersion forces but is too far away to be hydrogen-bonded directly to that water

molecule. In ice -Ih, the next-nearest neighbors, at the relatively large distance of 4.50 A,

are second neighbors as counted outward along the bond network. In the denser ice

structures, by contrast, the neighbors at 3 to 4 A are actually fourth or more distant neighbors

as counted along the bond network. This resuhs from a 'doubling back' feature of the

network, and allows the near neighbors to be accommodated with a minimum of bond

bending.

The bond-bending mechanism of densification is offset to some extent by a

lengthening of the hydrogen bonds in the dense ice phases, as shown by the mean bond

lengths listed in Table 1.2. The lengthening appears to be coupled to the bond bending and

reflects weakening of the bonds by bending. The weakening is shown also in infrared

spectra ofthe ice phases, the 0-H stretchingfrequenciesshowing a systematic increase as

the bonds lengthen [33].

Although hydrogen-bond bending is abundant in the dense forms of ice, the particular

pattern of bond bending represented by the original proposal of Bemal and Fowler [25] is

11

7^:^w^i^
not realized, because, as mentioned before, none ofthe ice phases is a structural analog of

quartz [26]. The failure of a quartz analog to occur resultsfromthe differences between the

bond-bending energetics of hydrogen bonds and Si-O-Si Hnkages, and from differences

between the packing requirements that arise when the relatively large oxygen atoms lie at

the tetrahedral centers, as in HjO, and when they lie at or near the centers of the bond

linkages, as in SiOj. For these reasons and because silicon is able to achieve 6-coordination

under pressure, the general structural analogy between SiOj and HjO is incomplete [34].

1.2.2.1. Ice-forms IL IIL V and IX

In the less dense high-pressure phases -II, -III, and -V, the network of hydrogen

bonds is a complete tetrahedralframework,like in ice -Ih and -Ic. Typical features ofthe

bent hydrogen-bond networks in the less-dense high-pressure phases are illustrated by ice

-V. In this structure there are 4-rings of water molecules, which are more compact groups

than the 6-rings that occur in ices -I and -Ic. The hydrogen-bond lengths in the 4-rings are

somewhat longer than in the rest oftheframework,so the 4-rings cannot be considered as

tightly bonded units. The 0 - 0 - 0 bond angles for a representative water molecule in ice -V

do not match the H-O-H angle of 104.5, they varyfrom87 to about 128. The hydrogen-

bond lengths generally increase over the length 2.75 A in ices -I and -Ic are evident. A

nonbonded near neighbor made possible by the tetrahedral distortion is the molecule at

distance 3.28 A.

The basic crystal structure of ice -II contains ice -I-like units buih out of puckered

'six-rings' of water molecules (Figure 1.2). These units are linked together in a more

12

yat^^^&j'Vt^St^
^ ^

Figure 1.2. Crystal Structure of Ice-II

13
compact way than in ice -I. The coordination ofthe oxygen atoms is markedly distorted

from the ideal tetrahedral angle of 109.5. Each oxygen atom has four nearest neighbors

at distances of 2.80 A. The distortion supports the existence of bent hydrogen bonds in ice

-II and a nonbonded near neighbor appears at 3.24 A from the central water molecule.

Ice -HI bears signs offtirtherdistortion from the ideal tetrahedral structure and

hydrogen bond lengths vary from 2.76 to 2.80 A. The structure can be described as

consisting of helical chains of hydrogen-bonded oxygen atoms held together by further

hydrogen bonds. Oxygen atoms can thus be members ofthe helix or act as links, bonding

four separate helical segments. Ice -IX has an ice -III structure, but with a neariy ordered

proton arrangement [35, 36].

1.2.2.2. Ice-forms VI. VH and VIH

In ices -VI, -VII and -VIII, a new structural feature appears, which is not present in

ice structures of lower density. This structural feature called as 'self-clathrate' structure

(because of its analogy with clathration in crystalline hydrates [37]) by Kamb [38], is built

by incorporation of two identical but independent tetrahedral frameworks. The two

independentframeworksinterpenetrate but are not interconnected by hydrogen-bonds. Each

frameworkfillsvoid space in the other, therefore the appropriate designation 'self-clathrate.'

Self-clathration is favored as a way of obtaining high packing densities of water molecules

without requiring excessive hydrogen-bond bending.

In ice-VII, which is the densest form of ice, the self-clathrate feature consists ofthe

interpenetration of twoframeworksof ice -Ic type. This arrangement places each water

14
molecule equidistant from eight neighbors that lie at the corners of a cube. The water

molecule at the center is hydrogen bonded to four of these, in perfect tetrahedral

coordination, and is nonbonded, but tetrahedrally arranged, with the other four. The

repulsion between each molecule and its four nonbonded nearest neighbors causes

lengthening ofthe bonds to 2.96 A. This 0 - 0 distance represents the longest and weakest

hydrogen bonds occurring in any ofthe ice phases. Because ofthe 'self-clathrate' structure,

ice -VII can return to the ideal tetrahedral bond geometry ofthe low-pressure ice phase Ic.

Bond-bending strain, in this case, essentially reduces to zero.

In ice -VI, bonded neighbor distances are 2.81 A and each oxygen has eight

nonbonded neighbors at 3.51 A. The interpenetration of two complex structures demands

a wide distribution of donor angles rangingfrom76 to 128.

In ice -Vm, a distortion occursfromthe structure of ice -VII, such that the hydrogen

bonds retain their length of 2.96 A, but the four nonbonded neighbors appear at two different

distances from the central molecule.

It is worth noting that the structural relationship between the various forms allow

crystallographic transformations over relatively low energy barriers among them [39]. These

transformations primarily take place by the bending [15], not the breaking, ofthe hydrogen

bonds. It is also important to note that, whereas the nearest-neighbor hydrogen-bonded

0 - 0 distances in all the ice forms are close to the same, between about 2.74 A and 2.95 A,

non-hydrogen-bonded second-neighbor distances are shortened from about 4.5 A in ice-Ih

and ice -Ic to 3.24 -3.51 A in ice -II, -HI, -V, -VI, and -IX, andftirtherto 2.95 A in the very

dense polymorphs, -VII and -VIII. Thus, the second-neighbor 4.5 A feature rather than the

15
often used 2.8 A feature acts as a more reliable 'finger print' for the presence ofthe open

tetrahedral bonding found in ice-Ih.

1.2.2.3. Amorphous Water

Besides these crystalline forms, low-density amorphous water (LDAW) and a high-

density amorphous water (HDAW) have attracted recent attention because of their possible

relationship whh the liquid [40-42]. The structure of these two amorphous forms merely

reflect the bonding properties ofthe various ice forms; LDAW has a near-neighbor structure

resembling that of ice-Ih or Ic, while the HDAW contains features found in the moderately

dense ice forms, -II, -III, -V and -VI.

1.3. Structural Models for Liquid Water

In order to 'understand water' in a molecular sense and to account for the results of

experimental observations made upon liquid water, one must resort to the consideration of

hypothetical models. The structure of water has been the subject of much experimental and

theoretical research for over a century. Since 1892, a very large number and variety of

models have been proposed with regard to the structure of liquid water, and some of them

will be described here.

1.3.1. Survey of Early Work

The various models proposed by numerous scientistsfromthe time of Rontgen until the

late sixties can be lumped into three main classes.

16

^mvs^

^ nT-XV^" ^ ^ ' .
1.3.1.1. Mixture Models

A mixture model describes liquid water as an equilibrium mixture of species that are

distinguishable in an instantaneous picture. This concept of liquid water has been

recognized for many years. As mentioned above, thefirstpublished statement to this effect

was that of Rontgen in 1892 [21]. In general, mixture models are discussed in terms ofthe

molecular species of water involved, the method of obtaining thermodynamic parameters,

and the partition functions appropriate to the model.

Most ofthe mixture models were introduced for special, and limited, purposese.g.,

that of Hall [43] to show how the anomalous acoustic absorption of water might be

accounted for. The elaborately quantitative model of Eucken [44] and the largely qualitative

model of Frank and Evans [45] were proposed, in part, to lay ground work for suggested

explanations of some striking properties of aqueous solutions. The mixture models of

Grjotheim and Krogh-Moe [46], Wada [47] and that of Davis and Litovitz [48] brought more

properties together, but none of these seem likely to be able to satisfy the criterion

established by the low-angle X-ray scattering in the matter of densityfluctuations[49]. The

shape ofthe low-angle scattering curve which, following Narten and Levy [49], places limits

on the kinds offluctuationsin density which can occur in the liquid.

The mixture model of Eyring et al. [50] is unique in applying to water a special

theory ofthe liquid state, the 'significant structure' theory [51]. The proposed two-state

model for water structure consists of an equilibrium mixture of clathrate-like clusters of

approximately 46 molecules dispersed in an ice-III-like structure. The ice-I-like clusters

exhibit the density of ice -I and are assumed to almost disappear by 4 C.

17
A simpler application ofthe significant structure method [51] pictures water as a

mixture of 5 kinds of species, each defined as a water molecule of which 0,1,2,3, or all 4 of

its bondingfiinctionsare actually bonded. A number of two-state mixture models have been

proposed based on this concept [52]. An elaborate treatment ofthe (0,1,2,3,4) mixture is that

of Nemethy and Scheraga [53], who proposed afive-statemixture model. They assumed that

the cooperative character of hydrogen bonding results in the formation and dissolution of

clusters of hydrogen-bonded molecules. These clusters are short-lived, forming and melting

as a consequence of local energy fluctuations. The clusters are considered to be embedded

in, and in equilibrium with, monomeric 'unbonded' water. The interior of the clusters

contains molecules with all four hydrogen bonds unbroken. Molecules with three, two, and

one hydrogen bonds can be found on the surface ofthe clusters. The molecules occupying

the space between the clusters have all four bonds broken. Unfortunately, there are now

known to be logical inconsistencies in this theory. One of these is an erroneous

combinatorial expression for the number of ways of distributing N molecules among the five

categories, and the assumed energetic interaction [54] of bonds involving the same

molecule.

K. Buijs and G. R. Choppin [55] investigated the near-infrared spectra of water and

resolved the broad absorption band in the 7700 to 9090 cm'* region into three components

located at 8000, 8330 and 8620 cm"^ They attributed these components to the existence of

three species of water molecules corresponding to two, one, or zero hydrogen bonds. The

mole fractions of the corresponding three species were calculated from the temperature

18

;W W.K ^yHj!''' ."TM?" >''"


dependence ofthe intensities ofthe three component bands. But no partitionftjnctionor

structural model was suggested.

Walrafen [56-58] made extensive measurements of the Raman spectra of the

intramolecular OH- and OD- stretching modes in water. In pure water [57], the Raman

intramolecular spectrum showed an asymmetric contour which is decomposed into four

Gaussian contributions. These contributions may be grouped into pairs: two intense low-

frequency components that decrease with temperature; and two weak high-frequency

components that increase with temperature. The temperature dependence ofthe high- and

low-frequencypairs of components leads to an isosbestic point, that is, a frequency at which

the intensity is independent of temperature. The occurrence of such an isosbestic point is

consistent with the existence of two species of water molecules. Walrafen has identified

these two species as water molecules that are hydrogen bonded with C2V symmetry and those

that are non hydrogen-bonded. Furthermore, Walrafen et al. [59] have compared the

stimulated Raman O-H stretching spectra in concentrated solutions of HDO in a mixture of

H2O and D2O. The results of this investigation provide strong evidence for the existence of

these distinguishable species in water. This evidence strongly favors the mixture model but

is in complete disagreement with the continuum model. More of this will be discussed in

Chapter VI.

1.3.1.2. Interstitial Models

Interstitial models are a special class of models: one species of water molecule is

supposed to form a hydrogen-bonded framework containing cavities in which the other

19
species, single, non-hydrogen-bonded water molecules reside. The idea that the increase in

density when ice melts arises from the invasion by interstitial molecules of some of the

empty spaces in the ice lattice was put forward by Samoilov [60]. From the radial

distribution curves found by Morgan and Warren [26], it was evident that liquid water has

a higher density of neighbors than ice at about 3.5 A. Samoilov also noted that each

molecule in ice -I is 3.47 A distant from six 'cavity centers.' These facts suggested that

liquid water is similar to ice but has molecules in the cavities.

Danford and Levy [61] and Narten et al. [62] made extensive calculations of radial

distribution functions for an interstitial water model similar to Samoilov's. The framework

in this model is an ice -I lattice that is permitted to expand anisotropically with increasing

temperature. The interstitial molecules are located in the cavities as in Samoilov's model but

the ratio of the interstitial to framework molecules is constrained to reproduce the

experimental density. By employing least-squares methods, the authors [62] were able to

calculate the radial distribution functions in close agreement with experiment. From their

studies, it was found that the occupancy of cavity sites increases from 45 percent at 4C to

57 percent at 200 C.

The somewhat different notion that liquid water could be regarded as a 'water

hydrate* was put forward by Pauling [63]. Pauling suggested that the configuration of

molecules in liquid water might resemble the clathrate compound, chlorine hydrate. In this

model, groups of 20 hydrogen- bonded water molecules form open, pentagonal

dodecahedra in which non hydrogen-bonded water molecules reside. The dodecahedra can

be packed together in ways that allow them to be linked by hydrogen bonds. If they are

20

*^S!iSS5"y^^T-.t'V^':">" "T-'^
packed as in chlorine hydrate, 46 molecules in each unit cell form the hydrogen-bonded

framework, and theframeworkencloses 8 cavities. Assuming that each cavity contains an

HjO monomer, the calculated density of this structure is 0.98 g/cm^ neariy that of liquid

water.

The thermodynamic implications of Pauling's model were explored by Frank and

Quist [64]. By allowing a variable degree of occupancy ofthe interstitial sites they found

they could account for the P-V-T properties up to 30 C and 2000 kg cm"^ Since the degree

of occupancy changes only slightly with temperature, the configurational heat capacity is

only about 0.55 cal/mol C, and thus the model does not account for the large heat capacity

of water. Frank and Quist concluded that, although structures of the type proposed by

Pauling may exist in water, the liquid cannot be composed entirely of framework and

interstitial molecules. They suggested that some fraction of the molecules must be in

transition between framework and interstices and that these transitional structures are

responsible for most ofthe configurational heat capacity.

1.3.1.3. Continuum Model

A continuum theory describes water as having essentially complete hydrogen

bonding, at least at low temperatures, but as having a distribution of angles, distances, and

bond energies. Whereas a mixture model implies different proportion of the species at

different pressures, a continuum model considers the bond energy to change with

temperature and pressure because of changes in the distribution of bond lengths and of

distortions ofthe angles.

21

*i*^<<'*?"'.f.^.r..' ' ^
Pople [65] established the modern continuum theory of water, following the first

structural theory work of Bemal and Fowler [25], providing an algebraic theory in which

the majority of hydrogen bonds between neighboring molecules are regarded as distorted or

bent; and it is the energy of this distortion that is to be determined, rather than the bonds to

be classified as broken or unbroken by some arbitrary criterion.

Pople showed how one might estimate the average angle of distortion of hydrogen

bonds in water. He assumed that the mutual orientation of two water molecules in the liquid

is determined only by the energy required to distort the hydrogen bond between them. He

described this energy by a 'hydrogen-bond bending-force constant.' Using classical statistics,

he also estimated the average angle between the 0-H bond direction and the O-O axis of

two hydrogen bonded neighbors.

Pople showed that when this formalism was used to calculate a radial distribution

function for the relative positions ofthe molecules, the result was consistent with what had

been deduced by Morgan and Warren [26]. Pople demonstrated that his model can account

for the dielectric constant of water and for the decrease in volume of ice -I upon melting.

Also, the model can account for the heat capacity and thermal energy of water.

Nemethy and Scheraga [53] have objected to Pople's model, arguing that a liquid so

extensively hydrogen-bonding would be highly viscous. Another objection to Pople's model

[66] was based on qualitative arguments to the effect that highly distorted hydrogen bonds

are unlikely in liquid water. Eisenberg and Kauzmann [67] believe that the' hydrogen-bond

bending-force constant' is too simple to give a good representation of the forces between

water molecules.

22

T-

wmssw^'^y^h^^T'Wm jfw^.-^m-^f^
>*r-

Bemal [68] had proposed a model for water based on his concept of a liquid as an

intrinsically irregular stmcture. Water molecules in the liquid, like those in the ices, are

considered to be four-coordinated; but the networks of linked molecules in the liquid are

depicted as irregular and varied, in contrast to the orderly networks of a few basic types that

are found in the ices. Bernal [68] believes that five-membered rings are a frequent

configuration in the liquid, but thatringscontaining four, six, seven, or even more molecules

are also part ofthe networks.

1.3.2. More Recent Models

Because of the rapid rate at which new data and insights about water started

becoming available, many ofthe detailed models (for fuller discussion ofthe older theories,

see [12, 53,69,70]) which have been proposed have been overtaken by events and are now

of hardly more than historical interest. There are essentially two routes along which the

theoretical study of water has been carried out during more recent times: Analytical and

Molecular Level approaches. A brief review of these methods is presented here, more

elaborate description of these approaches can be found in a book by Robinson et al. [71].

1.3.2.1. Analvtical Approach

Purely analytical approaches have the advantage of providing a mathematical

interpretation of the phenomena studied, sometimes giving formulas that can be used by

experimentalists to interpret their data.

23
1.3.2.1.1. Lattice and Cell Models More recently sophisticated statistical theories

have been developed which incorporate molecular orientation and directional bonds between

molecules into a simple liquid model in which the molecular centers occupy certain sites of

a lattice [72-74].

An alternative route for describing water uses cell models or free-volume theories

[75], which also postulate a body centered cubic structure and treat the liquid as a lattice

gas. These theories, however, assume that the field acting on each molecule is rapidly

fluctuating, and therefore, may be replaced by an average field of spherical symmetry. Both

the lattice model and the cell model are based on a simple configurational description of

local molecular order at a dynamically uncorrected level, and have an attractive appeal [16].

In addition to the lattice and cell theories, a number of other quantitative theoretical

studies have emerged. Rice and his coworkers have applied the random network models to

liquid water and amorphous solid water [76]. Belch and Rice have presented a critical

overview of these type of theories with the aid of computer simulation methods [76].

1.3.2.1.2. Using Integral Equations. Theoretically, the stmctural properties of two-

component systems can be investigated through analytical or numerical solutions of integral

equations such as the Percus-Yevick equation [77] and the hypernetted-chain (HNC)

equation [78-80]. All these formalisms can be derived from the well-known Omstein-

Zemike (OZ) equation [81]. These integral equations can be solved numerically by iteration

methods or analytically with the aid of mean spherical approximations [82] or perturbation

theories [83]. The liquid properties described by these theories agree only crudely with the

24
MD results, although the model has been greatly improved by Ichiye and Haymet [84] using

a more accurate scheme to solve the integral equations.

One semiempirical approach is the reference interaction site model (RISM) [85]

which solves the Ornstein-Zemike (OZ) type equations [81]. These types of models

represent the interactions between molecular entities by a sum of pairwise additive

spherically symmetric potentials between sites located in the molecules. RISM type models

were extended by Rossky and coworkers [86,87] for the purpose of studying water.

Angell [88] and Speedy [89] have suggested a number of theoretical avenues of

approach to the liquid water problem through their experimental studies of amorphous solid

water, the supercooled liquid and the 'singular temperature' near 227.4 K.

1.3.2.2. Molecular Level Approaches

The birth of the contemporary computer simulation technique began a new era of

theoretical studies of water. This technique provides a direct route for converting the

microscopic level information of a system to macroscopic properties of interest.

Approximations enter at the more fundamental level of intermolecular interaction potentials.

Energies, structure, etc., of liquid water then emerge automatically from the computation,

rather than through any ad hoc assumptions.

There arefivemost commonly used computer simulation methods. They are Monte

Carlo [90,91], Molecular dynamics [71], Stochastic dynamics [90,91], Energy minimization

[92] and Normal mode analysis [5].

25

i^mmgmaimmsm^
T.iiiiirt*^

Among these computer simulation techniques, Monte Cario and Molecular Dynamics

provide the most promising approaches for water [71]. Unlike the Monte Cario approach,

which is essentially a probabilistic form of computer simulations, molecular dynamics is a

deterministic approach. (A complete description of molecular dynamics simulations on water

is given in ref [71].) They contain the least number of approximations, can be applied to a

variety of experimental conditions, are readily adaptable to many statistical ensembles, and

provide detailed molecular-level insights about the liquid structure that cannot be extracted

from other methods.

The most apparent disadvantage of computer simulation approaches is the limitation

of sample size. Generally, the simulation cell contains no more than several hundred to

several thousand particles. With contemporary computer hardware and sophisticated

software, it is possible to extend the sample size to perhaps a million particles. Any

successful simulation technique requires the development of reasonable potential models,

and, in the case of water, these models must intrinsically be rather complicated, greatly

reducing the sample sizes that may be employed.

1.4. Interfacial Water

Water in very small volumes plays a dominant role as the medium that controls

structure, function, dynamics and thermodynamics near biological membranes or in other

confined regions of space. The study of interfacial water has received extensive attention

during the past decade because ofthe role it plays in many physical, chemical and biological

processes. A knowledge of its interfacial properties is of great importance for a full

26

r"?Si":."?V:JJ-i
JS'WSiTTvS
understanding of the stability and behavior of proteins, membranes and various naturally

occurring homogeneous substances. However, its highly complex nature at interfaces has

prevented a good understanding of its physical properties when it is confined within very

small spaces [93].

Since the structures within liquid water are so fragile with respect to temperature and

pressure changes, it would be expected that effects similar to those in the bulk liquid might

be even more perturbed by surfaces. If the surface is chemically and physically inert, the

inability ofthe water molecules to extend their hydrogen bond structure into such a surface

could have the same effect as a temperature increase. If the surface interacts with water, the

ability of those molecules forcibly oriented near the surface, to maintain normal hydrogen

bonding with their neighbors creates the same type of perturbation. The open ice-I-type

tetrahedral structure would have to give way to a denser bonding, of ice-II-type with more

tightly packed hexagonal channels combined with 4- and 5-membered ring structures.

The study of liquid water near various surfaces or interfaces has been performed

using computer simulation methods. The system of surfaces considered includes a layer of

liquid water confined between two solid surfaces [94,95], liquid-filled pores [96], the

liquid/vapor interface [97,98], the ice/water interface [99], a self-supporting thin film [100],

ionic interfaces [101], the water/nonpolar solid interface [102] and strong laser fields [103].

Microscopic quantities of interest in the interface include the density profile, surface tension,

molecular orientational and translational order, the rate of molecular diffusion and

thermodynamic quantities such as heat capacity, compressibility and expansitivity. The

27

i'Y.^^.'Z^T'^'^^m
results of these simulations indicate that the properties of liquid water at a surface or

interface are generally quite different from its bulk properties.

Unlike theoretical work, very few detailed experimental investigations of interfacial

water have been carried out. It has proven difficuh to elucidate the liquid water stmcture

near a general surface. Most ofthe experimental work concerning the stmcture of interfacial

water is derived from nuclear magnetic resonance spectroscopy [104,105], ultrafast laser

spectroscopy [106,107], neutron scattering [108] and x-ray diffraction [109] studies ofthe

confined water phases. Direct measurements by ultrafast laser experiments have been

performed by Robinson et al. [106] in order to study dynamic properties such as diffusion

of perturbed water in various interfaces and confined volumes. Experimental studies using

optical second harmonic generation methods have been used to probe the stmcture and the

orientational order ofthe vapor/water interface [110,111].

Recently, experimental studies concerning reverse micelle stmctures were conducted

in our laboratory [112]. A reasonably well-studied medium, an amphiphilic molecule

Aerosol-OT (AOT) was used to investigate the properties of water within a confined volume.

Water encapsulated in these systems is thought to mimic water close to biological

membranes or proteins [113]. These reverse micelle systems are often used as models of

water pockets because of their well-defined size, shape and aggregation number. Results of

this investigation using lifetime studies indicate that the interfacial water in these systems

is orientationally more stmctured than the bulk water.

28
1.5. Present Work

There seems to exist today a general feeling that a complete understanding of the

properties of liquid water is unfathomable and in any case cannot be found through mixture

model concepts. The basic validity of mixture model theories has been questioned [114] and

the ideas on which this model is based have been discredited by Kauzmann [114] and some

others [115]. It may be largely due to the fact that essential data, especially experimental

data in the supercooled region, were not then available. The purpose ofthe present work is

to try to get the venerable [21] mixture model for liquid water back onto its feet. As more

experimental data accumulate, it becomes more and more difficult to ignore the accurate

quantitative assessments ofthe properties of this liquid that can be obtained from this simple

model.

The mixture model proposed here is based on the concept, promoted by Kamb [15]

from his extensive crystallographic work on the ice polymorphs, that intermolecular

hydrogen bonds in the liquid can be bent without being broken. The melting of ice-Ih can

be interpreted in terms of a material having both the normal open hydrogen bond stmcture,

representing a low density, and bent bonds that reflect a higher density. The 'bent bond' idea

agrees with the opinions expressed by Pople [65], some years eariier.

Following these ideas, a semiempirical theory, attempting quantitatively to correlate

thermodynamic and dynamic effects of bulk liquid water with the various properties ofthe

ice polymorphs was developed in our laboratory [41,116-119]. This theory is a

modernization of the so-called 'mixture' or 'two-state' models of the liquid. Instead of

broken' and 'intact' hydrogen bonds [53,120], or other hydrogen bonding arrangements [121]

29
comprising the two states, one 'state' has an open ice-Ih local tetrahedral bonding

arrangement, while the other is considered to be a mixture of bonding types prominent in the

higher density forms of ice. Particulariy considered were ice -II (lacking the ordered proton

arrangement in the crystal), ice -HI, ice -V, ice -VI. In all of these dense stmctures, the

hydrogen bonds are intact in the first neighbor-shell, but are bent in the second-neighbor

shell. These higher density stmctures exist in the liquid at low temperatures [15] and grow

in with increasing temperature in accord with experimental findings [122,123], which have

found an increase in the second-neighbor 0 0 stmcture near 3.4 A at the expense ofthe

second-neighbor -4.5 A stmcture prominent in ice -Ih. These ideas about the nature ofthe

liquid mentioned above would explain the large local densityfluctuationsin water that have

been observed experimentally [124] and in computer simulations.

Through our analyses [116,117], it was discovered that remarkably good correlations

can be found, from the deep supercooled regime to temperatures approaching the boiling

point, that are consistent with the density maximum of liquid H2O and liquid D2O.

Extensions of this work have been able to interpret quantitatively the temperature-dependent

isothermal compressibility [118] including the minimum near 46.5 C, and has very recently

led [119] to a very simple 'thermal scaling law' (similar to the one mentioned for the

viscosities [125]), to about four decimal place accuracy, connecting the densities of H2O and

D2O. Extensive work [41,116-119] in these areas indicates that all of these features of liquid

water can be quantitatively understood using the model outlined in the beginning ofthe next

chapter.

30

P'A-^a^ii^
A premiere set of experimental data on water, the set quantitatively most amenable

to a precision description because of its high level of accuracy and extensive thermodynamic

range, is the liquid density. The density of liquid water will thus constitute the main focus

of Chapter H. In Chapter HI, the attention shifts to the isotope effect on the density anomaly.

The simple 'temperature-scaling phenomenon' is discussed in relation to the densities of

isotopic water in Chapter FV. Chapter V deals with pressure effects, including the isothermal

compressibility and Chapter VI gives a brief account ofthe experimental results that provide

strong support for the model proposed here. The ultrafast laser technique mentioned in the

previous section has been extended to study liquid water confined between two extremely

flat quartz plates and interfacial water confined in a thin film supported by a gold wire ring.

The results are presented in chapter VII. The conclusions of this work are enumerated in

Chapter Vni.

31

ims^S^^^T^r.s'r;^.
CHAPTER II

THE DENSITY ANOMALY OF LIQUID H2O EXPLAINED

BY THE MIXTURE-MODEL

The primary motivation for the initial formulation of "mixture models" [21] was to

offer an explanation for the density maximum. In the mixture model for liquid water

proposed here, there are supposed to be two or more general types of intermolecular bonding

configurations, a bulky stmcture having a low density and a dense stmcture. The bulky

bonding form has a low density such as occurs in ice -Ih, and the dense bonding form has

a stmcture similar to the thermodynamically most stable dense forms of ice, e.g., ice -II, -III,

and -V [ see Table 1.1].

This model is not to be conceived as a 'mixture of ices' having even modest long-

range order, but rather is a rapidlyfluctuatingmixture of intermolecular bonding types found

in the most stable polymorphs of ice. As the temperature or pressure is raised, the capacious

bonding weakens, and the stmcture fluctuates more and more between the diverse, but

energetically similar [10] types of intermolecular bonding.

Increasing temperature (and pressure) also causes the intermolecular potential

surfaces in liquid water to cooperatively soften andflattenout, an effect that is very likely

related to the Pauling-Fowler model [126,127] for cooperative librational melting in solids.

The potential surfaceflatteningaccompanying the breakdown of the capacious stmcture,

givesriseto a greaterfluidityfor the dense stmctures, contrary to ordinary intuition. At high

temperatures (or high pressures), where more of the higher density, less viscous type of

32
bonding is already present, the dynamic properties of water begin to behave more normally

with variations in T (or P).

The expression for the pressure/temperature-dependent specific volume, V(P,T), of

liquid water in the mixture model proposed here is:

V(P.Tl = ffP,T)VfJ^^T) . f,{P,T)V^fP,T) (2.1)

where fj and fn represent the mass fractions ofthe capacious and dense components as a

function of pressure and temperature. As there is no ice polymorph that has a density

between those ofthe low density forms: 0.92 g cm '^ (ice -Ih and ice -Ic) and those ofthe

most stable - high density forms: 1.18 (ice-II), 1.15 (ice-III), 1.26 (ice-V), and 1.34 gcm'^

(ice-VI); it will be assumed that fi + fn = 1. This assumption does not mle out the

presence of a variety of stmctures as long as such stmctures can be grouped according to

their local densities into either type I or type IL The fraction fn might be expected to

approach zero at supercooled temperatures and to increase with increasing T or P. For the

above mathematical expression to be convincing, both forms I and H must have normal

positive thermal expansion coefficients. Thus, the specific volumes ofthe forms increase

as the temperature is raised.

33
2.1. Oualitative Description nf the Density Anomaly

The density anomaly, i.e., the density increase with increasing temperature in the

range -35 to +4 C, giving a density maximum at +4 C, can be explained using the above

equation in view ofthe 'proportionality effect.'

Taking the derivative of equation (1) with respect to P we get the expression given

{dviBT), = [{v^^ - v;) dfjdT ^//aF/az) ^/.(dv^/dT)], (2.2)

in eq 2.2. Term 1 on the RHS of equation (2.2) is negative, since V, > VQ, and the fraction

fn of dense stmcture increases with increasing T. Terms 2 and 3 have positive signs because

of normal thermal expansion.

At low temperatures, the magnitude of term 1 dominates the magnitudes ofthe other

two terms, because ofthe steep increase of fn with increasing T. Therefore, thefirstterm

gives a dominating negative contribution to the volume changes at these temperatures. So,

it is the strongly increasing 'proportionality' of fn that gives rise to the initial decrease of

volume or increase of density with increasing temperature and thus to the density maximum

at +4 C.

Thus, the density of liquid water at supercooled temperatures anomalously increases

with an increase in temperature because of the 'proportionality' effect. However, at

temperatures above +4 C, the proportion of term 2 becomes larger and the increase of fn

'saturates'. The (afn/aT)p part of term 1 no longer increases and the magnitude of the

negative term 1 diminishes. At higher temperatures, the positive terms 2 and 3 begin to

34
dominate, and the liquid water behaves like a 'normal' liquid with density increasing \^ith

temperature.

2.2. Experimental Density Data

The aim here is to fit to a good precision the available density data and to compare

the output parameters with the bonding characteristics ofthe low and high dense forms of

ice.

Liquid water can exist at atmospheric pressurefromabout -40 C to about 325 C,

although, density data are available for only a part of this range [18]. Early exploration of

water properties in the supercooled region was that of Despretz [128] who in the 19th

century measured accurately the density of water down to -9 C. Not much progress was

made in the area of expansion of water in the metastable region for several decades since

Despretz's determination. It was only in 1912 that Mohler extended the density

measurements to -12 C [129] by using a two-meter-long capillary sample container, but no

further penetration of the supercooled range was undertaken for another fifty years.

However, during the period 1965-69 several workers [130-132] determined the density

versus temperature relation for water at temperatures down to -34 C, using fine capillaries

as sample containers. For a very long time, the accuracy ofthe density measurement in the

supercooled region has been known to be better than three decimal places. However, most

recently. Hare and Sorensen [133] measured the density of liquid water up to four decimal

point accuracy from 0 to -34 C.

35
Measurements on liquid water below -10 C are somewhat diflicult to perform

because ofthe likelihood of ice formation in samples of normal size and cleanliness. Hence,

it is necessary to study very small samples in which the probability of freezing is sharply

reduced. The onset of crystallization during cooling is a highly unpredictable and sample-

dependent event. Much effort has been expended on attempts to understand the causes of,

and the limits to supercooling of water [134]. The choice of 'small sample' technique

depends largely on the type of measurement being performed. The use of capillary

techniques has been the method of choice for several studies of density in relation to

temperature and pressure. Using fine columns or capillaries, very small samples can be

isolated in a manner such that all volume changes are registered by one-dimensional

movement of a well-defined boundary, the meniscus.

Although the measurement is simple in principle, involving only the measurement

ofthe length ofthe water column, systematic errors in the measurements occur due to sample

size effects. The early measurements of Mohler employed capillaries whose diameters

ranged from 188 - 380 pm, so those data should present minimal size effects. Comparison

of Mower's data with the single data mn reported by Schufle [130] using a 133 ^m diameter

capillary shows that the two sets of data agree within (1 -2) x 10 g cm" . Schufle's density

data values (extended to -21 C), compared with the data of Zhelezyni and Hare-Sorensen,

are seen to be a bit too small at the lower temperatures.

Leyendekkers and Hunter [135] have argued that the effect of surface energy on the

free energy of small samples can be considerable, especially for samples in the range of 2

pm to 90 pm. Comparison of Hare-Sorensen's data [136] obtained using 300 pm diameter

36

fc'MiE^^ '.'''.^'.\L^^'if^
capillaries to eariier data obtained from measurements in smaller capillaries [131-133]

showed that the smaller capillary data, most often was larger than the measurements taken

using the 300 pm diameter capillaries. This excess density in smaller capillaries was found

to be inversely proportional to the capillary inside diameter. Hare-Sorensen's 300 ^im

capillaries were large enough to avoid this excess density and hence yielded bulk water

density data. All capillary density data show larger densities than Hare-Sorensen's 300 pm

data except those of Zheleznyi [132] at very low temperatures. Zheleznyi's data show

smaller densities for temperatures less than -25 C for some unexplained reason.

Kell carried out empirical fits ofthe experimental density datafrom0 to 100 C and

showed that some rationalfiinctionsgave a good and efficient representation ofthe data [18].

Kell recommended for ordinary water, at atmospheric pressure, a Pade approximant [137]

containing seven adjustable parameters

p = {GQ + a^t + cq'^ + af + a^t'^ ^ a^t^ ) I { \ + bt). (2.3)

The calculations were performed using double precision on an IBM 360/67 and the

coefficients were determined by least square fittings.

The output values of this equation for 133 temperatures between -30 and 150 C are

listed in Table III in ref [18], The number of decimal places provided by this author is six

from -30 to -1 C andfrom+101 to +150 C and seven decimal places for the data between

0 and 100 C. These tabulated data provide probably one to two decimal point precision

beyond the actual experimental accuracy. Therefore, for easy comparison with experimental

37
data, thefinaldecimal point in the Kell data will be rounded off in this presentation, though

it was retained in all ourfittingprocedures.

This compilation currently provides the premiere values for the density of liquid

water. When Kell'sfittingequation is extrapolated below 0 C, the agreement with the most

recently measured densities of Hare and Sorensen [136] in the supercooled region are

apparently much better than Kell realized [138]. Kell densities [18] in this important

temperature range are compared in Table 2.1 with published experimental density data [130,

132,136]. Kell's analytic equation is smoothed, so, when using it one need not consider any

experimental uncertainty or 'noise.' Between -30 and 0 C, Kell's equation agrees within one

part in 10'* with the experimental densities of Hare and Sorensen. For this reason, we select

the Kell [18] smoothed extrapolated data and the Hare-Sorensen 300 pm experimental data

[136 ] as the most reliable throughout the supercooled regime.

2.3. Density Analysis

Precise quantitative fits of the temperature-dependent density data for liquid HjO

at atmospheric pressure can now be carried out with the experimental data and the concepts

discussed in the previous sections. Analytical forms for the pressure/temperature-dependent

quantities Vi(P,T),Vn(P,T), and fn(P,T) in terms of a set of temperature-independent fitting

parameters at fixed at 1 atm pressure are required in order to perform the density analysis.

Vi and Vn are expanded in their standard forms and are then inserted in eq 2.1 to be

utilized for the analysis. Standard forms for V, and Vn are:

38

Mv' *l*,ifflLai(,li h ';. Llxe

"^^^^^^^T^^^^
Table 2.1 Liquid Water Densities (g cm'^) at 1 atm from Kell's Equation Compared \sith
Experimental Supercooled Densities

t,C Schufle" Zheleznyi'' H-S' Kell''


-34 0.9751 0.9775
-32 0.9793 0.9809
-30 0.9829 0.9839 0.983 85
-25 0.9895 0.9895 0.989 59
-20 0.993 36 0.9936 0.9935 0.993 55
-15 0.996 16 0.9964 0.9963 0.996 28
-10 0.997 96 0.9983 0.9982 0.998 12
-9 0.998 25 0.9986 0.9984 0.998 40
-8 0.998 51 0.9989 0.9987 0.998 65

-7 0.998 75 0.9991 0.9989 0.998 87

-6 0.998 98 0.9993 0.9991 0.999 08

-5 0.999 23 0.9995 0.9993 0.999 26

-4 0.999 38 0.9996 0.9994 0.999 41

-3 0.999 53 0.9997 0.9996 0.999 55

-2 0.999 67 0.9998 0.9997 0.999 67

-1 0.999 77 0.9998 0.9998 0.999 76

0 0.999 85 0.9999 0.9999 0.999 840 1

39
Table 2.1. Continued

t,C Schufle* Zheleznyi'' H-S^ Kell'


+4 1.0000 1.0000 0.999 972
+10 0.999 700
+15 0.999 100

+20 0.998 204

+25 0.997 045

+30 0.995 647

+40 0.992 216

" DatafromReference 130.

'' DatafromReference 132.

" DatafromReference 136.

^ DatafromReference 18.

40
VfFJ) = VfPJ^ [1 . afP) {T-T^. p/P) {T - T^^ . ...] (2.4)

VjiiP^T) = Vij(PJo) [1 - ^jfP) (T-T^. p^/P) {T - 7i)^ . ...] (2.5)

Introducing eqs.2.4 and 2.5 into eq.2.1 gives us an equation with seven adjustable

parameters: V,(P,To), Vo(P,To), a,(P), an (P), Pi(P), Pn(P) and To(P).

Choosing a suitable form for the proportionality factor fn was a time consuming and

tedious process, because ofthe properties expected ofthe parameter from the discussions in

sections 2.1 and 2.2. Several forms of fn were tried for this purpose. Some ofthe forms that

were rejected for one reason or the other are tabulated and presented in Table 2.2. The form

chosen for our analysis is given in eq (2.6),

A{T-T^^B(T-T^'
fjj(P,T) = D tanh (2.6)
1 ^C{T-T^

This mathematical form has the desired temperature behavior, approaching zero as T-> TQ

and reaching a constant value, D ^ 1, as the temperature is raised. This form is certainly not

unique and in fact is somewhat arbitrary. It is the simplest of other successfijl forms that

gave a six- to seven-decimal-point precision required for the density fits. The pressure-

dependent parameters A, B, and C in themselves have little or no physical significance. All

of the physical significance resides in the quantity fn (P,T). Equation (2.6) thus

41

.^/^ 'Arm*, ii'il f J r l


Table 2.2. A Few Forms of f,(T) Tested for the Density Analysis

No. Formoffi(T) Parameter Chisquare


Remarks
To = 200.0 K(f) 1.2 X IQ-
//T) = exp [-B ( r - n n
B +ve, n=0.9

To= 185.0 K 1.7 X 10"^


- =[Ua(r-r(,) ^b\T-T^'] a & b +ve
/,('/)

To= 150.9 K 1.8x10 -3


= [1+ b\T-T^^\
A{T) b +ve

To= 228.0 K(f) 1.4 X 10"^


C & D +ve
''o 'o
B-ve, n = 28

To= 233.1 K 6.7x10-


For {T-T^ < 50, //T) = \-A {-j^f A&B+ve

For {T-T^ > 50 / / 7 ) = Aexp {-BT) A & B +ve 1.0x10-

Combining Eqs (5) and (6); To= 195.0 K 1.5x10-3


T-T T-T B A,B,& p+ve

42
introduces four more adjustable parameters, giving in all 11 adjustable parameters for each

pressure P, four more than in the Kell smoothing ftinction.

It is obvious that the number of adjustable parameters, even considering the six-to

seven-decimal-point accuracy of the smoothed density data, is too large. Any fitting

procedure involving such a large set of parameters would be subject to errors, particulariy

the smaller a, p, and B parameters, because of mutual compensation effects. In many

preliminary and test fits it was found that these fits attempt to match the meaningless zeros

beyond the reported number of significant figures in the experimental data. Thus, it will be

necessary to find ways of reducing the number of adjustable parameters without in any way

forcing a fit toward the proposed mixture model.

The parameter D reflects the 'degeneracy' of dense component configurations. If

there is only one capacious component I, having open tetrahedral bonding, one dense

component 11, and no other stmctural forms, D would be equal to 1/2, since the populations

ofthe dense and capacious components would become equal at large T. However, this is

very likely not the case for water, since there are expected to be at least four distinct (i.e.,

having different densities) higher-density bonding forms in the liquid, corresponding to the

bonding configurations of some ofthe most stable ice polymorphs. Probably then, D should

be closer to 0.8.

As the temperature rises, all the specific bonding forms must give way more and

more to fully disordered stmcture. In fact, the best preliminary fits up to 40 C always

showed D to be very close to unity, which, taken literally, would mean that the open

tetrahedral bonding tends to disappear entirely at high temperatures in lieu of more densely

43
packed and disordered stmctures. It was therefore elected to set D equal to unity in the fits

However, it was found that, for all practical purposes, it really does not matter too much

whether D has a value of 1/2, 1, or any value in between, since the A, B, and C parameters

were found to compensate, so that the resulting physically significant fn values for a given

temperature always turned out to be the same. This factftirtherillustrates that the form of

fn is immaterial to the fitting procedure, providing, of course, that the chosen form has the

right mathematical properties. It is the value of fn itself as aftinctionof temperature (and

pressure) that really matters. Eq.2.6 is just one of many forms that could have been chosen

to express the temperature-dependent fn in terms of a set of temperature-independent fitting

parameters. Thus, excluding D, there still remain a total of 10 adjustable parameters for the

density fits.

Through a large number of preliminary data fits, it was found that the parameter TQ

almost always had a value between 224 and 230 K. In the present mixture model

description, this temperature is the temperature at which all intermolecular bonding in liquid

water resemble that in disordered or amorphous ice-Ih or -Ic [139]. Very interestingly, these

TQSfromthe density fits are not farfromthe 'singular temperature', 228 K, found by Angell

and co-workers [140] from a variety of independent considerations. Of further interest are

the output values of the two volumes, Vj(P,To) and Vn(P,To), together with their thermal

expansion coefficients. The values of fn(P,T), through the parameters A, B, C, and TQ, are

also interesting, since this quantity should have connections [41] with modern experimental

data in the supercooled regime: radial distributionfianctions[122,123], small-angle x-ray

44

*I^T^-^- \ " '*


scattering [141], and activation energies [142] for dynamical processes as aftinctionof P and

T.

Thus, it is seen that the output V's, a's, and P's, in addition to fn and To, but not the

particular mathematical form of fn, will have physical significance for the density of liquid

H2O discussed in this chapter and for the density of liquid D2O to be discussed in chapter

III. The pressure dependence of these parameters will be required for discussions ofthe

isothermal compressibilities in chapter V. Since only atmospheric pressure is being

considered in chapters II through IV, the pressure dependence of all the variables will be

suppressed in these chapters.

2.4. Densitv Fits

Among the approximately sixty fits employed, only seven of them are reported here

in detail:

Fit 2,1. an exact low-temperature (^ +4 C) algebraic fit of eight selected density

data points to determine the eight parameters A, B, C, To, Vi(To), Vn(To), a,, and an, with

D=l and the P's set equal to zero for this very low-temperature regime;

Fit 2.2. an exact algebraic fit, like fit (2.1) but extended up to +40 C, of eight

selected density data points and the P's were set equal to zero;

Fit 2.3. a least-squares fit of 37 points between -30 and +22 C; during the density

analysis, the density and the thermal expansion coefficient are fixed at the ice-Ih values at

225 K for the form I component;

45

n^FTT;
Fit 2.4. a least-squares fit of 39 data points from -30 to +40 C (the number of data

points being equally balanced above and below +4 C) to attempt to refine, for this high

temperature regime, the values ofthe four basic volume parameters - Vi(To), Vn(To), a,, and

an - when the A, B, C, and To parameters, and thus the values of fn(T), are fixed at their fit

(2.1) output values;

Fit 2.5. a least-squares fit exactly like fit (2.3) but including the p parameters to

check if the density and the a parameters change to accommodate the additional parameter;

Fit 2.6. a least-squaresfitsimilar to fit (2.3), but with the A, B, C, and To parameters,

notfromfit(2.1), but deduced from temperature-dependent activation energies for dynamic

processes as outlined in ref [41]; and

Fit 2.7. a least-squares fit, with the samefixedfit(2.3) parameters, of 39 data points

between -30 and +70 C to determine best values of all six volume parameters (V's, a's, P's)

when a higher temperature regime is considered.

Also, a briefer description of some ofthe other fits (fits 2.8-2.11) examined is included in

the following section.

2.4.1. Exact Algebraic Fits

2.4.1.1.Fit(2.n

Of the sixfitsmentioned above,fit(2.1) is perhaps the most important, as it provides

unequivocal connections between the density data and the mixture model described by eq

(2.1). Also, this fit is very interesting, since included are only low-temperature data. In this

46
msism

temperature regime, the model is expected to be the most accurate because the multiplicity

of intermolecular bonding stmctures is minimized.

For this fit, only data from the supercooled regime up to +4 C was used. This

temperature range is one that provides the most rapidly increasing value of fQ(T). Eight

uniformly spaced data points in this temperature range were chosen to create eight required

equations, and then solved algebraically for the eight parameters in equations 2.1,2.4-2.6,

accepting only the physically relevant solution from the nonlinear equation set. Specifically,

the Kell [18] density data at -30, -25, -20, -15, -10, 15, 0, and +4 C were chosen. Such an

algebraic fit cannot be 'forced' in any way, since eight density points are to be fit exactly by

equations containing eight unknowns, with nofreedomfor manipulation or choice of the

output parameters.

The approach to a unique set of correct output parameters is only possible from

'noiseless' experimental density data, providing there is at least one-decimal-point accuracy

in the densities beyond that desired for parameters. If the experimental data to be fit have

errors of a few percent, it would be expected that parameter variations of at least that

magnitude can occur without greatly modifying the 'exact' algebraic solution. Since Kell's

density data arefroma smoothing equation, only round-off 'noise' in thefinaldecimal place

is present. Recent experimental density data of Hare-Sorensen in the supercooled regime

can be used to calibrate Kell's extrapolated data to the required level of accuracy.

Thus, the procedure chosen here should give accurate low-temperature liquid

parameters for comparison with the density and the thermal expansion coefficients ofthe ice

polymorphs, and also with the other properties of water and ice. The resulting high precision

47
parameters just 'drop out' of such a fit, and the calculated densities, of course, match exactly

the six-to seven-decimal-point density precision given by Kell for these eight data points in

this temperature range. The resuhing parameters for this fit are given in Table 2.4. The near

agreement ofthe output's with ice-Ih and ice-II densities is noteworthy along with fact that

To is close but somewhat lower than the 'singular temperature' at 228 K [140]. Recalculated

densities from these fit (2.1) parameters for temperatures between -30 and +40 C are

compared with the smoothed Kell [18] density data in Table 2.3. The agreement remains

excellent for other temperatures in and somewhat beyond +4 C as well and the deviations

at higher temperatures are in the expected direction for the missing p terms.

2.4.1.2. Fit (2.2)

This fit, like fit (2.1), again finds an exact algebraic solution using just eight data

points but this time the analysis is extended to +40 C. Kell's densities at -30, -25, -15, -5,

+4, +15, +25, and +40 C were selected, so as to give a balance between data points above

and below +4 C. Some ofthe output parameters in this case were pi(To) = 0.919 g cm'^,
-3
Pii(To)= 1-215 gem , a, = 2.61x10-* K-', and an = 1.50x10''K-\ A list of the all output

parameters obtained from fit (2.2) are listed in Table 2.4. The only significant difference in

these parameters and those obtained from fit (2.1) is in the pi(To) value, which shows a

modest increase toward the density of ice-V and -VI as more high-temperature density data

are included in the analysis. The precision achieved in fits (2.1) and (2.2) extends somewhat

48

w^rr?-^^^
Table 2.3. Liquid Water Densities (g cm"^) at 1 atm from Kell's Equation Compared with Fit
(2.1) Densities

t,C Fit (2.1)" Kell'' (Fit (2.1)-Kell)


xlO"'
-34 0.977 36
-32 0.980 87
-30 0.983 85 0.983 85 0.0
-25 0.989 59 0.989 59 0.0
-20 0.993 55 0.993 55 0.0
-15 0.996 28 0.996 28 0.0
-10 0.998 12 0.998 12 0.0
-9 0.998 40 0.998 40 0.0
-8 0.998 65 0.998 65 0.0

-7 0.998 87 0.998 87 0.0

-6 0.999 08 0.999 08 0.0

-5 0.999 26 0.999 26 0.0

-4 0.999 41 0.999 41 0.0

-3 0.999 55 0.999 55 0.0

-2 0.999 67 0.999 67 0.0

-1 0.999 76 0.999 76 0.0

0 0.999 840 0.999 840 0.0

49
Table 2.3. Continued

t,C Fit (2.1)* Kell" (Fit (2.1)-Kell)


X 10-'
+4 0.999 972 0.999 972 0.0
+10 0.999 700 0.999 700 0.0

+15 0.999 100 0.999 100 0.0

+20 0.998 206 0.998 204 0.2

+25 0.997 050 0.997 045 0.5

+30 0.995 659 0.995 647 1.2

+40 0.992 255 0.992 216 3.9

* Recalculated Densities from the Exact Algebraic Fit.


'' DatafromReference 18.

50

-Ml

^^ftp^BS^
Table 2.4. Fitting Parameters from Fits (2.1) and (2.2)'

Parameters Fit (2.1) Fit (2.2)


A 4.54866 E-2'' 4.10847 E-2
B 3.6522 E-4 2.7983 E-4
C 8.69196 E-2 8.82543 E-2

To 225.334 225.337
Vi(To) 1.08761 1.08801

Pi(To) 0.9195 0.9191

Vn(To) 0.84632 0.822748

Pn(To) 1.1816 1.2154


a, 4.78005 E-4 2.61161 E-4
ttn 1.29473 E-3 1.49824 E-3

PI 0.0 0.0

Pn 0.0 0.0

* The units are as given in the text.

Fit (2.1) : Exact algebraic fit of 8 Kell densities between -30 and +4C;

Fit (2.2): Exact algebraic fit of 8 Kell densities between -30 and +40C.

"Read as 4.54866x10-1

51
2.4.2. Least-Squares Fits

A least-squares fitting procedure (see the Appendix) is employed when more

experimental points than the number of parameters are considered. A number of such fits

were performed. A concern with any fitting procedure of this type is that Kell's data set

contains many more density points above the temperature of maximum density than below

it. Therefore, the dominating high temperature data could distort the more interesting

limiting low temperature parameters.

To attempt to avoid any such problem, a 'blind'fitwasfirstcarried out with only 37

densities between -30 and +22 C, eighteen below 4 C and eighteen above 4 C in Kell's

table. A least-squares fit of all the eight fitting parameters to this limited set of data

indicated again that TQ was close to 228 K, and also that Vi(To) and Vn(To) were in the

expected ranges for an ice-Ih/dense-ice mixed bonding model. For this number of fitting

parameters, one more than Kell's, and for such a limited range of data, fits better than

warranted by the experimental data could be obtained over wider ranges of the parameter

values than was thought realistic. For example, V,(To) ranged from 1.05 to 1.10 cm' g'\ while

Vn(To) ranged from 0.80 to 0.85 cm'g-'. With all thefittingparameters variable, even the

six-to seven-decimal-point agreement with the Kell data is an insufficient criterion for

obtaining unique values for the parameters, particulariy for minor parameters a's, p's and

B. The thermal expansion coefficients a,, and an changed more radically, between about

5x10-' and 1x10-' K'\ and were sometimes even negative, as these two quantities attempted

to compensate for variations in the two volumes.

52

i^.^-^-i:'.ti^'"--'"'^'-
2.4.2.1. Fit (2.3^

To avoid the errancies, discussed in the previous section, the number of adjustable

parameters had to be reduced. If the density, i.e., V,(To), and the thermal expansion

coefficient a, are fixed at the ice-Ih values (at 228 K), -0.9234 gem'^ and -1.31x10-" K-\

respectively [20], then all of Kell's values could be fit within 2 parts in 10'for the 37 data

points between -30 and +22 C. The x^ error (<1 Q-'^) for the fits was far better than justified

by the experimental uncertainty. To make the density p, ofthe ice-Ih component consistent

with the slightly lower To being found from fit (2.1) and fit (2.2), the density for fit (2.3) was

fixed at 0.9237 g cm "^ instead of 0.9234 g cm '^ This modification gave rise to less than

a 3% variation in the other variable parameters. As shown in Table 2.5, the output

parameters for this fit are. To = 225.815 K, p, = 0.9237 g cm "^ (fixed), pn= 1.1877 g cm "^

(cf ice-n, p'-1.18 gcm"^at238K), ai= 1.31xl0-^K-' (fixed), and an = 1.40515x10-'K-\

an being larger than aj, as perhaps expected for the more compact component. Even for

these relatively low temperatures, it is probable that the a parameters are somewhat

distorted, because ofthe missing higher-order thermal expansion terms, Pi and Pn- The

choice of Vi(To) and a, in this fit goes against thefindingsfromfits(2.1) and (2.2), which

definitely imply that the type-I stmcture is somewhat less compactly bonded in the liquid

than in crystalline ice-Ih, and that its thermal expansion coefficient is larger. However, the

value of To = 225.815 K was about the same as that obtained for fit (2.1), as was pn(To) =

1.1877 gem -3

53
Table 2.5. Fitting Parameters from Fit (2.3)'

Parameters Fit (2.3)


A 4.04834 E-2"
B 2.5963 E-4
C 8.33749 E-2
To 225.815
V,(To) 1.08262"=
P,(To) 0.9237*=
Vn(To) 0.841990

Pn(To) 1.1877
! 1.31 -4"=
an 1.40515 E-3

Pi 0.0"=

Pn 0.0"=

" The units are as given in the text.

"Read as 4.04834 xlO-l

*= Parameter fixed.

54

f^^K .^'<rW.l}Ml.. W
The conclusion derivedfromcomparing the output parameters for fits (2.1) and (2 2)

with those from this fit is that Kell's densities can be reproduced, to better than the

experimental uncertainty, from the two-state mixture model using a variety of fitting

methods. The least-squares output parameters from these close density fits, however, vary

to some extent as they attempt to compensate for one another's variations. The principal

parameters do not vary (relatively) as much as the minor parameters. During the fittin

procedures, it became clear that, with even a lesser number offittingparameters, it should

be possible to fit the equation to the density data over a wider temperature range with good

precision.

2.4.2.2. Fit (2.4)

Therefore, a decision had to be made as to which of the fitting parameters were

expendable, while still retaining the desired physical content in the fits. The density and

thermal expansion coefficient of ice-Ih are known to much lower precision than required to

fit the liquid data accurately, and , in any case, as mentioned just above, may not be entirely

appropriate for the liquid environment anyway. In a number of these 'blind' fits, it was also

found that, even though the A, B, and C parameters varied a good deal, the resulting values

of fn(T) were pretty stable and agreed within better than 10% with the values obtained from

thefit(2.1) parameters. Thus, it was decided that, in order to reduce the variable parameters

to a manageable number without distorting much their physical significance. A, B, C, and

To could be fixed in the following least-squares fits.

55
This least-squares fit of Kell's density data for temperatures ^+ 40 C employs just

four adjustable parameters, setting the p's to zero, D to unity, andfixingthe A. B, C, and To

parameters at their fit (2.1) output values. The analysis utilizes 39 of Kell's density data

points in the range -30 to +40 C. Only every fifth data point above +20C is selected in

order to give an approximately equal number of data points above and below the temperature

of maximum density. The output parameters from fit (2.4) are listed in Table 2.6.

2.4.2.3. Fit (2.5)

Thisfithas been extended to include the p parameters, so that compensation among

the values V, a, and p parameters could be evaluated. In this case, the six output parameters

have values as follows: Pi(To) = 0.9197 g cm "^ pn(To) = 1.1953 g cm *^ ai= 9.34 x lO"'*

K-\ and an = 6.06 x IQ-^R-', p, = 5.34 xlQ-^R-^, PQ= 2.42 xlO-^K-^(see Table 2.6 for a list

of all parameters obtained). It is noticed that the p, (TQ) value hardly changed at all, still

accurately matching the density found for fits (2.1) and (2.4). The density Pn(To) has moved

only very slightly toward ice-V and -VI densities, thus fairiy well preserving the ice-Ih/ice-II

mixture model picture. However, a, has doubled, while an has diminished by a factor of 2

to compensate for the presence ofthe nonzero P values. The exact reason for these latter

changes is not known but could have to do only with the increased uncertainty in the minor

parameters when a great number of adjustable parameters is employed.

56
Table 2.6. Fitting Parameters from Fits (2.4) and (2.5)*

Parameters Fit (2.4) Fit (2.5)


A 4.54866 -2""= 4.54866 -2'
B 3.6522 -4"= 3.6522 -4"=
C 8.69196 -2"= 8.69196 E-2'
To 225.334' 225.334'
Vi(To) 1.08739 1.08801
Pi(To) 0.9196 0.9197
Vn(To) 0.84745 0.822748

Pn(To) 1.1800 1.1953


a, 4.57558 -4 9.34 -4
an 1.29374 -3 6.06 -4

Pi 0.0"= 5.34 -6

Pn 0.0"= 2.42 -6

" The units are as given in the text.

Fit (2.4) : Least-squares fit of 39 Kell densities between -30 and +40C, without the P's;

Fit (2.5): Least-squares fit of 39 Kell densities between -30 and +40C, with the P's
included.

"Read as 4.54866 xlO-l

'Parameter fixed.

57
2.4.2.4. Fit (2.6)

It was demonstrated by Bassez, Lee and Robinson [41] in great detail, how a large

number of thermodynamic and transport properties of liquid water, supercooled to

superheated, can be interrelated through empirically determined temperature-dependent

activation barriers. In their paper it was suggested that the height of these barriers is a

quantitative measure ofthe departure ofthe local ice-I-like stmcture, i.e., f,(T) = AH(T)/

AH(To), the ratio of activation energies at T and To. It was found that both the densities and

dynamic data could be fit to rather high temperatures with the mixture model, though the

precision demanded was less by a few percent than what was achieved in the current study

In any case, it was feU that it may be worthwhile to follow the lead provided by Bassez et

al.[41] to see if a procedure similar to the one there can provide accurate densities. A

measure of self-consistency with the dynamic data would be retained in this way, and a

better extension of all these data to high temperatures may be possible in the future.

For the present purposes, the temperature dependent dynamic rates from ref [41 ]

were smoothed, and AH(To) was chosen to be 14,820 J mol"'. A preliminary least-squares

fit to eq 2.6 ofthe dynamic f,(T) values (from the ratio ofthe activation energies at T and To)

[41] in the temperature range 238-423 K wasfirstperformed to obtain A, B, C, and To. The

To value found was 224.8 K, again somewhat lower than the Angell [140] singular

temperature of 228 K, but in pretty good agreement with Tofromfit(2.1). Since these values

are so close, it was decided to fix the To value for the preliminary fit at 225.334 K, the

temperature found in fit (2.1). Corrected A, B, and C parameters were then deduced from

eq 2.6 with thisfixedtemperature using least squares. These A, B, C parameters (see Table

58
2.7) were then used in the density fits, only V,(To), Vn(To), a and an being varied. The

same 39 densities from fit (2.4) were utilized for this fit and all the output parameters are

given in Table 2.7.

2.4.2.5. Fit (2.7)

This fit extends the density analysis to +70 C. For both ofthe fits, fits (2.1) and fit

(2.4), the p values were set to zero. When including high temperature density values,

setting the P values to zero would undoubtedly distort the values ofthe a parameters and

could affect the values of V,(To) and Vn(To) to some extent as well. A fit ofthe density data

to +70 C certainly requires the inclusion of p in eqs 2.4 and 2.5. Thus, for fit (2.7), the

parameters A, B, C, D, and To were fixed, as was done for fits (2.4) and (2.5), but allow the

other six parameters to vary. The Kell data points are again balanced above and below

+4 C, with a total of 39 points being used. The resulting output parametersfromthis fit are

listed in Table 2.8.

The primary analyses, presented here, considered Kell's data only over the

temperature range -30 to +70 C. Densities for temperatures above +70 C were also

examined. The high-temperature mixture model fits were found to give adequate agreement

with the density data, but the physically significant quantities begin to vary as more high-

temperature data are included in the fits. The greatest sphere of reliability ofthe model, and

thus its main interest should reside in the low-temperature regions anyway, where not too

59
Table 2.7. Fitting Parameters from Fit (2.6)'

Parameters Fit (2.6)


A 4.52912 E-2"''''
B 4.3486 -4''"*
C 9.56974 -2''"*

To 225.334'
Vi(To) 1.09180

Pi(To) 0.9159

Vn(To) 0.81340

Pn(To) 1.2294
a. 8.37014 E-4

an 1.49776 -3

Pi 0.0'

Pn 0.0'

' The units are as given in the text.

Read as 4.52912x10-2

Parameter fixed.

''Parameter from dynamic data.

60
Table 2.8. Fitting Parameters from Fit (2.7)'

Parameters Fit (2.7)


A 4.54866 -2"-'
B 3.6522 -4'
C 8.69196 -2'

To 225.334'
V.(To) 1.08717

Pi(To) 0.9198

Vn(To) 0.83369

Pn(To) 1.1995
a. 1.09085 -3

"n 3.39868 -3

Pi 7.4987 -6

Pn 3.2862 -6

' The units are as given in the text.

"Read as 4.54866x10-1

'Parameter fixed.

61

.mV:iS^<'im6mii^^
many intermolecular bonding forms play a role and whereftillydisordered stmcture is at a

minimum. In spite of such problems, the major parameters - V,(To),Vn(To), TQ, and fo - did

not vary more than about 6% in a variety of fits tried for temperatures up to 100 C, so a

remnant ofthe mixture model must exist at fairly high temperatures.

A comparison with the Kell data ofthe recalculated densities in the temperature

range -34 to +100 C from the least-squares fits (2.4), (2.6) and (2.7) is provided in Table

2.9, and Table 2.10 compares some f,(T) values obtained from the dynamical data with those

derived from the densities of fit (2.1).

2.5. Other Fits

Some other fits were performed to determine how far from the ice-Ih or ice-II

densities could the related Vi(To) and Vn(To) parameters stray and still provide good fits to

the density data. From the exact algebraic fit (2.1), it becomes fairly obvious that, by fixing

one parameter, seven are adjustable and thus seven data points can be fit exactly, no matter

what value orfixedparameter was given. Furthermore, with this number of parameters, the

same number used by Kell [18], least-squares fits might also give excellent agreement with

the experimental data. Thefitspresented in Tables 2.3-2.8 constitute a necessary condition

for the validity of this type of model. The question now being posed is whether those fits

provide a sufficient condition and 'strong support' for a mixture model.

The parameters Vi(To) and Vn(To) cannot take on absurdly different values than we

have derived. Otherwise, nonphysical values ofthe other parameters would arise: negative

a's, but much more seriously, negative V, negative fjor fn, or negative TQ. TO evaluate this

62
Table 2.9. Recalculated Densities for Least-Squares Fits (2.4), (2.6) and (2.7) Compared
with those from Kell Data*

t.C Kell'" Fit (2.4) Fit (2.6) Fit(2.7)


-30 0.983 85 +1 -1 0
-25 0.989 59 -1 +1 -1
-20 0.993 55 -1 +1 0
-15 0.996 28 0 +1 0
-10 0.998 12 0 0 0
0 0.999 840 +1 -4 -1
+4 0.999 972 +2 -2 0
+10 0.999 700 0 +1 -1
+20 0.998 204 -3 +4 0
+30 0.995 647 -4 +1 +1
+40 0.992 216 +7 -7 0
+60 0.983 199 +135 +5 -1
+80 0.971 798 +547 +174 +8

+100 0.958 364 +1457 +705 +71

" Shown in columns 3-5 are density differences in thefinaltabulated decimal place between
the Kell data and the recalculated densities using the parameters from the three fits
mentioned.

For t < 0 C, +2 means that the recalculated value is greater than the Kell's by two in the
fifth decimal place.

For t ^ 0 C, the difference is in the sixth decimal place.

63

y^mm^mm
Table 2.10. Temperature-Dependent Fractions of Ice-I-Type Bonding from Fit (2.1),
compared with those from Fit (2.6) from Temperature-Dependent Dynamic Data

t, c Fit (2.1) Fit (2.6)


-34 0.6930 0.7039
-30 0.6517 0.6642
-20 0.5757 0.5897
-10 0.5199 0.5336
0 0.4746 0.4871
+4 0.4585 0.4703
+20 0.4014 0.4104
+40 0.3422 0.3475

+60 0.2925 0.2945

+80 0.2501 0.2494

+100 0.2137 0.2108

64

^fS^^V:/^^--^^!
point, four algebraic seven-parameter data fits were carried out for the same temperature

range -30 to +4 C, as fit (2.1), where one ofthe V(To) values was fixed, D = 1, and the p

values were set equal to zero. The fixed V value in each of these fits was set either 10%

higher or 10% lower than the corresponding value in fit (2.1). The resuhsfromthese fits are

listed, in brief, in Tables 2.11 and 2.12.

Looking at the results in these two tables, it is evident that the output parameters

obtained do not make much physical sense. Fit (2.8) corresponds to an unexpectedly high

Pn(To), trying to compensate for the fixed low Pi(To). The parameters obtained from fit

(2.9), because of the trend toward similar values of Vi(To) and Vij(To), remained totally

unphysical, with negative To, and f^T) values always close to zero, which lead to an unstable

aj. The parameters from fit (2.10) and fit (2.11) look at least physically reasonable except

that aj is negative in both cases. In addition to the less than pleasing parametersfromthese

four fits, the densities above the fitting range are not so well represented as for fit (2.1).

Recalculated densities from the three sets of physically acceptable parameters indicate that

they are off*fromKell's densities by 2 or 3 in the fourth decimal place. For example,

recalculated densities at +40C (36 C above the fitting range) for fits (2.8), (2.10) and

(2.11) are 0.992 13, 0.992 36, and 0.992 08, respectively. These values may be compared

with the fit (2.1) density at this temperature of 0.992 26 and the Kell density of 0.992 22.

65
Table 2.11. Fitting Parameters from Fit (2.8) and Fit (2.9)'

Parameter Fit (2.8) Fit (2.9)


d
To 219.96

V,(To) 1.20289" 0.98418"

P,(To) 0.831" 1.016"

Vn(To) 0.49206 0.93096

Pu(To) 2.032 1.074


d
a, 5.56 -5'
d
otn 4.84 -3
d
fn (-^4 C) 0.36

* The units are as given in the text.

Fit (2.8): Vi(To) fixed at a value which is 10% higher than the corresponding value in fit
(2.1);

Fit (2.9): V,(To)fixedat a value which is 10% lower than the corresponding value in fit (2.1).

" Parameter fixed.

'Read as 5.56xlO^

** Parameters unphysical.

66

i^^i^mi^mss^iis^/^' ftxiv-
Table 2.12. Fitting Parameters from Fit (2.10) and Fit (2.11)"

Parameter Fit (2.10) Fit (2.11)

To 229.67 231.51
Vi(To) 1.05442 1.04572

P.(To) 0.948 0.956

Vn(To) 0.94067" 0.76964"

Pu(To) 1.063" 1.299"

a, -4.72 -4' -9.98 -4

an 8.25 -4 6.84 E-3

fn (-^4 C) 0.58 0.16

" The units are as given in the text.

Fit (2.10): Vn(To) fixed at a value which is 10% higher than the corresponding value in fit
(2.1);

Fit (2.11): Vn(To) fixed at a value which is 10% lower than the corresponding value in fit
(2.1).

" Parameter fixed.

'Read as-4.72x10-^.

67
2.6. Discussion

The main purpose ofthe present work was not just tofitdensity data accurately This

has already been done by Kell [18] and others. The purpose was to check the consistenc\

ofthe derived mixture model parameters with solid-state bonding parameters, and eventually

with other type of experimental results, and thus to create an alternate way of thinking about

liquid water when experiments are being designed and analyzed.

The exact agreement in Table 2.3 between the calculated mixture-model densities

using fit (2.1) parameters and actual HjO densities in the temperature range -34 to nearly

+40 C, in addition to the fact that the output parameters in Table 2.4 have close

relationships with bonding characteristics ofthe open and dense forms of ice, indicates that

a necessary condition for the validity of the mixture model has been achieved. The near

uniqueness ofthe fitting parameters, discoveredfromthe examination of a large number of

fits, goes part way in providing a sufficient condition for such validity. In chapters that

follow, through consistency checks for other types of experimental data on liquid water,

further convincing arguments supporting the validity ofthe model have been created.

Near the lowest experimentally available temperature of supercooling, -34 C (the

homogeneous nucleation temperature), Table 2.10 indicates that thefractionf, of ice-I-type

bonding from the fit (2.1) analysis is -0.69. At +4 C thisfractiondrops to -0.46. The

intermixture of these very large proportions of material having such great stmctural and

density parity should surely be detectable experimentally. In Chapter VI, several

experiments that lend important but independent support to the presence of two temperature-

dependent bonding forms ofthe types expectedfromthe density analysis are discussed.

68
J
The somewhat weaker ability ofthe equations to reproduce the density data into the

higher temperature regimes is not caused, as can be inferred from Table 2.9, by the form of

f, but rather it is caused by the expected breakdown of eqs.2.1, 2.4 and 2.5 at high

temperatures, where a plethora of bonding types occur. This bonding multiplicity has been

one ofthe main objections in the past to 'mixture models' of liquid water. Because ofthe

results presented in this paper, it is now felt that such objections lose much of their validity

at sufficiently low temperatures. Water is a completely different type of liquid than liquid

argon, on which most theories of the liquid state have been based and for which the

objections have more validity. Therefore, the mixture model concept, which is certainly

simpler than other proposed 'liquid theory' models, should be able to form a sound low-

temperature base for extending the understanding of liquid water to other thermodynamic

regimes.

69

. ^?:^mmamm^'^
CHAPTER III

ACCURATE MIXTURE-MODEL DENSITIES FOR D,0

In the present chapter, the main attention concerns the isotope effect on the density

anomaly. The density maximum in D2O occurs at a temperature (11.18 C) over 7 C

higher than in HjO (3.94 C). It is believed that the major factor contributing to this isotope

effect is that the parameter fi(P,T) is sensitive to intermolecular frequencies and zero-point

energies in the temperature-dependent potentials [41], causing the open stmcture in HjO to

break up more easily at a given temperature than that in DjO. Thus, the magnitude of

fj(D20) is expected to be greater than fi(H20) at all temperatures, pushing the density

maximum to a higher temperature in D2O. The isotope effects on the librations, which are

closely related to activation energies, are also responsible for isotope effects on the viscosity,

heat capacity, and the relaxation times [41]. The mixture model proposed in chapter II for

liquid H2O to explain its density anomaly, is now applied to the densities of liquid D2O in

the temperature range -20 to +100 C. The aim here is to obtain, through eq.2.1, precise

quantitative fits of the temperature-dependent density data of D2O and to compare the

results with those obtained for HjO.

3.1. Experimental Densitv Data

Some fairly recent experimental work on supercooled DjO has been performed by

Hare and Sorensen [133] on samples confined in glass capillaries of 25 pm diameter. These

density values, in the temperature range from -20 to +40 C, are presented in Table 3.1.

70
However, these data should not be taken at face value because of systematic errors arising

from surface energy effects [135]. As mentioned in the previous chapter, the effect of

surface energy on the free energy of samples in small capillaries can be considerable.

Because of this, the measured apparent density increases with decreasing sample size below

about 100 pm.

The density at atmospheric pressure of liquid HjO from -30 to +150 C is well

represented by Kell's rationalftinctionwith seven parameters [18]. A similarfijnctionwith

one less parameter and one lessfigureof accuracy was employed by Kell [17] to represent

the density of D2O from 0 to +101 C at atmospheric pressure. Using the coefficients for

the rationalfianctiongiven in Table 3 of ref [17], the densities for temperatures below 0 C

were calculated. These data, withfive-decimal-pointprecision, from -20 to +100 C are

presented in Table 3.1.

Comparison of Kell's data with Hare and Sorensen's datafrom0 to +40 C shows that

the agreement is fairly good, with deviations only of the order of 1-2 x IQ-^ g cm"^.

However, at temperatures lower than about -5 C, larger deviations are evident, with the

Hare and Sorensen density values being greater than Kell's. This is what would be expected

from surface energy effects. If a surface energy correction were applied, these two sets of

density data in the supercooled regime would no doubt be in much better agreement. For

example, comparing Hare and Sorensen's density data for H2O in capillaries of 25 and 300

pm diameters [133, 136], scaling the temperature to the same H2O/D2O (T- TQ) values (see

below), and applying the H2O error ratio to the D2O data gives almost perfect agreement

71
Table 3.1. Liquid D2O Densities (g cm-^) at 1 atm from Kell's Equation Compared with
Expenmental H-S Supercooled Densities

t,C H-S' Kell"


-20 1.092 40 1.091 81
-19 1.093 62 1.092 99
-18 1.094 80 1.094 09
-17 1.095 75 1.095 11
-15 1.097 52 1.096 96
-10 1.100 94 1.100 58
-5 1.103 33 1.103 07
0 1.104 79 1.104 69
10 1.106 06 1.105 99
11.18 1.106 00
20 1.105 28 1.105 34
30 1.103 19 1.103 24
40 1.100 01 1.099 96
50 1.095 70
60 1.090 60

70 1.084 75

80 1.078 24

90 1.071 12

100 1.063 46

"Reference 133 with linear interpolation.

"Reference 17.

72
with the Kell D2O extrapolated density data. For the present purposes, Kell's smoothed dati

will be used [17] to represent the experimental DjO densities for least-squares fit analysis

3.2. Densitv Fits

In order to fit the experimental density data of DjO using eq.2.1, analytical forms

for Vj(P,T),V(P,T), and f,j(P,T) in terms of a set offittingparameters are required. The

forms for V, and Vn used are given in eqs.2.4 and 2.5. In these equations, To is the

temperature at which fj(P,T) = 1, i.e., where all the bonding in the Hquid is represented by

the low-density, open hydrogen-bonded network work stmcture of heavy ice-Ih. However,

in the liquid state, this stmcture is expected to be highly disordered, similar to that in

amorphous solid water [143]. As stressed in the previous chapter, the proposed model,

because of missing long-range order, is by no means to be conceived as a 'mixture of ices'.

The parameters a, and an in eqs.2.4 and 2.5 are the linear, and p, and Pn are the quadratic,

thermal expansion coefficients. The arbitrary analytical function chosen for fn(P,T), given

by eq.2.6, bearing in mind that its value lies in the range 0-1. Substituting eqs.2.4-2.6 into

eq.2.1 gives a function with 10 adjustable parameters for each pressure P (with parameter,

D, being set equal to 1).

A number of fits were employed including some preliminary ones, but only two of

them are reported here in detail:

73
Fit 3.1. an exact low-temperature (^ +11.18 C) algebraic fit of seven selected density data

points to determine the seven parameters A, B, C, To, Vn(To), a and an, with V,(To)

fixed and the P's set equal to zero for this very low-temperature regime;

Fit 3.2. a least-squares fit of 13 data points from -20 to +70 C (the number of data points

being equally balanced above and below +11.18 C) to attempt to refine, for this high

temperature regime, the values ofthe six volume parameters - V,(To), V(To), a a,

PI and Pu- when the A, B, C, and TQ parameters, and thus the values of fu(T), are

fixed at their fit (3.1) output values.

3.2.1. Preliminary Fits

Through the use of the procedure given in the Appendix, some preliminary least-

squares fitting was performed, with the P's equal to zero, using all temperature points in

Table 3.1from-20 to +11.18 C. Through these fits it was found that the parameter TQ was

about 232 K, which is close to the singular temperature of 236 K for D2O given by Angell

[144]. However, such fits, as would be expected because of small number of adjustable

parameters for so few data points, gave unstable values ofthe output parameters V, and V^.

Hence, it was decided to fix the value of Vi(To), which should correspond roughly

to the D2O ice-Ih density at 232 K. In the eariier work described in Chapter II, this

correspondence was not exact. The density ofthe open ice-Ih-type bonding stmcture in the

liquid, because of missing long-range order, could be slightly lower than in the crystal at the

same temperature. To incorporate this correction, the ratio, 1.0222/0.9213, ofthe actual

densities [145] of D2O ice-Ih at 232 K [To(D20)] and HjO ice-Ih at 225 K [To(H20)] was

74
multiplied by the density, 0.9195 g cm^ ofthe ice-Ih component (at 225 K) used earlier for

the liquid H2O fits. Whh this procedure, the fixed value of V,(To) to be used in the liquid

D2Ofittingprocedure is 0.98020 cm^ g'^ This corresponds to p = 1.0202 g cm'^ compared

with the actual density [145], p = 1.0222 g cm-^ of D2O ice-Ih at 232 K.

3.2.2. Exact Algebraic Fh

3.2.2.1. Fit (3.n

Similar to the procedure described in section 2.4.1, an algebraic fit was performed

by utilizing seven roughly equally spaced points between -20 and +11.18 C. This creates

the seven required equations to solve algebraically for the seven output parameters (p, fixed

and P's = 0). The density values obtainedfromthis fit exactly match the five-decimal-point

density precision given by the Kell equation, not only for the data points used in the fit but

also for others out to neariy +30 C. The resulting parameters and density valuesfromfit

(3.1) are given in Tables 3.2 and 3.3.

3.2.3. Least-Squares Fit

3.2.3.1. Fit (3.2)

This fit was performed to include the higher temperature density data. In fit (3.2),

the parameters A, B, C, and To are fixed at their output valuesfromthe previous fit, and the

P's are included. This fit utilizes 13 of Kell's density values from -20 to +70 C, with an

equal number of density points above and below the density maximum. The analysis is

carried out by varying the six remaining adjustable parameters V,, VQ, a,, an, P,, and PQ.

75
The output parameters so obtained are given in Table 3.2, and the resulting density values

are found in Table 3.3.

A comparison with the Kell data ofthe recalculated densities in the temperature

range -20 to +100 C from fits (3.1) and (3.2) is provided in Table 3.4. Also, Table 3.5

compares f,(T) values obtained from H2O fit (2.1) density analysis with those derived from

D2O fit (3.1) density analysis in the temperature range -20 to +100 C.

3.3. Discussion

The densities obtained throughfit(3.1) agree very well with the experimental density

data, even beyond the fitting range. With the addition of the p parameters in the least-

squares fit, it is seen that the calculated densities are excellent out to nearly 100 C.

Considering the number of fitting parameters used in the model, the attainment of such

agreement is in itself not very surprising and Kell's fitting function for D2O contains only

six parameters. So, like in the case of liquid H2O density analysis, the concern is not

whether the density data can be fit to a high precision by the proposed mixture model, but

rather, the concern is whether thefittingparameters make sense.

Firstly, the density ratios of D2O and H2O have to be analyzed more closely.

Ignoring the zero-point effects and the often unreported variations in isotopic compositions

in the experiments, the density ratio of D20^^/H20^^ would be close to the molecular weight

ratio of 1.1117. As mentioned in the previous section , mainly because ofthe higher value

of To(D20), the required ratiofromthe actual ice-Ih crystal densities [145] is closer to

76
Table 3.2. Fitting Parameters from Fhs (3.1) and (3.2)'

Parameters Fit (3.1) Fit (3.2)


A 3.97129 -2" 3.97129 -2'
B 3.0258 -4 3.0258 -4'
C 7.34949 E-2 7.34949 -2'
To 231.832 231.832'
V,(To) 0.98020'-'' 0.98007

Pi(To) 1.0202 1.0203


Vn(To) 0.76854 0.75357

Pn(To) 1.3012 1.3270


ai 4.00995 -4 1.04376 -3

an 1.30045 -3 4.47931 -4

Pi 0.0' 6.39610 -6

Pn 0.0' 3.14581 -6

* The units are as given in the text.

Fit (3.1): Exact algebraic fit of seven D2O densities between -20 and +11.18 C.

Fit (3.2): Least-squares fit of thirteen D2O densities between -20 and +70 C.

"Read as 3.97129x10-2

' Parameter fixed.

See text.

77
Table 3.3. Liquid D2O Densities (g cm-') at 1 atm from Kell's Equation Compared with Fit
(3.1) and Fit (3.2) Densities

t.C Kell'' Fit (3.1) Fit (3.2)


-20 1.091 81 1.091 81 1.091 81
-19 1.092 99 1.092 99 1.092 99
-18 1.094 09 1.094 09 1.094 09
-17 1.095 11 1.095 11 1.095 11
-15 1.096 96 1.096 96 1.096 96
-10 1.100 58 1.100 58 1.100 58
-5 1.103 07 1.103 07 1.103 07
0 1.104 69 1.104 69 1.104 69
10 1.105 99 1.105 99 1.105 99

11.18 1.106 00 1.106 00 1.106 00


20 1.105 34 1.105 34 1.105 34

30 1.103 24 1.103 25 1.103 24

40 1.099 96 1.100 02 1.099 96

50 1.095 70 1.095 85 1.095 70

60 1.090 60 1.090 90 1.090 60

70 1.084 75 1.085 29 1.084 75

80 1.078 24 1.079 13 1.078 24

90 1.071 12 1.072 48 1.071 11

100 1.063 46 1.065 43 1.063 43

Fit (3.1): Exact algebraic fit of seven D2O densities between -20 and +11.18 C.

Fit (3.2): Least-squares fit of thirteen D2O densities between -20 and +70 C.

78
Table 3.4. Recalculated Densities for Fits (3.1) and (3.2) Compared with those from Kell'

t,C Kell Fit (3.1) Fit (3.2)


-20 1.091 81 0 0
-19 1.092 99 0 0
-18 1.094 09 0 0
-17 1.095 11 0 0
-15 1.096 96 0 0
-10 1.100 58 0 0
-5 1.103 07 0 0
0 1.104 69 0 0
10 1.105 99 0 0

11.18 1.106 00 0 0

20 1.105 34 0 0

30 1.103 24 +1 0

40 1.099 96 +6 0

50 1.095 70 +15 0

60 1.090 60 +30 0

70 1.084 75 +54 0

80 1.078 24 +89 0

90 1.071 12 +136 -1

100 1.063 46 +197 -3

' Shown in columns 3 and 4 are density differences in the final tabulated decimal place
between the Kell data and the recalculated densities using the parameters from the two fits
mentioned.

For example, t=30 C, +1 means that the recalculated value is greater than Kell's by one in
the fifth decimal place.

79

^ f.^^mmatmsmmm:
Table 3.5. Temperature-Dependent Fractions of Ice-I-Type Bonding from H2O Fit (2.1),
Compared with those from D2O Fit (3.1)

t. C H 2 O Fit (2.1) D20Fit(3.1)


-20 0.5757 0.6343

10 0.5199 0.5646

0 0.4746 0.5112

+4 0.4585 0.4927

+11 0.4321 0.4631

+20 0.4014 0.4291

+40 0.3422 0.3650

+60 0.2925 0.3120

+80 0.2501 0.2671

+100 0.2137 0.2287

80

^^^ v9.-\. '.-wvys


1.1095. In fact, the ratio of p,(H20)/p,(D20) for fit (3.1) has been fixed at exactly this

value. From Tables 2.4 and 3.2, the Pnratio is calculated to be 1.1012, which is somewhat

lower than the p, ratio. The ratios pn/p, for H2O (fit (2.1)) and D2O (fit (3.1)) are 1.2850

and 1.2754, respectively. The ratios are close, as expected, since only the mass, not the

lowest order bonding stmcture, is changed by the isotopic substitution. Thus, it is seen by

fixing the value of p,(D20) in fit (3.1), reasonable values are obtained for p(D20). More

about the relationship between isotopic density ratios will be discussed in the next chapter.

As seen from Tables 3.3 and 3.4, when higher temperature data are included in the

fits, like infit(3.2), there is a tendency for Pn(To) to tend toward the density of denser solid

forms (ice-V and ice-VI), and when P's are included, the a's change to accommodate them.

The same effects were observed for H2O when densities at higher temperatures were in the

density fits.

The fi values from the A, B, C and To parameters using eq 2.6 for the two isotopic

modifications at different temperatures was calculated (see Table 3.5). As per expectation,

the magnitude of fi(D20) is greater than fi(H20) at all temperatures. The f, values of H2O

and D2O at their temperatures of maximum density differ by only about +0.01, which is

within their uncertainty. This reveals that, when temperatures are scaled according to the

temperatures of maximum density (or to be more exact, their TQ values), the bonding

fractions in liquid D2O are not very different from those in liquid H2O. This is very much

in agreement with the observation [145] that the difference between H2O and D2O ice-Ih

forms is hardly significant, both with respect to the hydrogen bond lengths and the

tetrahedral angle.

81

^^^h:^ --^
3 i/'

As stated in the liquid H2O analysis in the previous chapter, the excellent density fits

that can be obtained from the ice-Ih/ dense-ice mixed bonding model for liquid water

certainly provide a necessary condition for the reality ofthe model. Further support for this

model is obtained from a number of experimental observations which are discussed chapter

VI.

82

RBBPy ^-.. ',V .7' i-aifei


CHAPTER IV

A SIMPLE RELATIONSHIP BETWEEN THE DENSITIES AND

OTHER PROPERTIES OF ISOTOPIC WATER

In the discussion section ofthe previous chapter, it was mentioned that the bonding

fractions in liquid D2O are not very different from those in liquid H2O. Therefore, it was

conjectured that if temperatures of H2O and D2O are scaled according to their temperatures

of maximum denshies, then the stmctural properties of these two liquids should be very

closely the same. According to this concept, shifting the temperature scale for H2O upward

by the difference between the temperatures of maximum density [17] of H2O (3.984 C,

where p = 0.999972 g cm'^ ) and D2O (11.185 C, where p = 1.10600 g cm'^), or about

7.2, would cause the stmctural properties of these two liquids to be nearly identical over

a fairly wide range of temperatures.

The simplest property of water for testing such a relationship is the density. In this

chapter, the vaUdity of this idea is tested for the densities within the four decimal point

experimental precision ofthe D2O density data.

4.1 Comparison ofthe H2O and D2O Densities


Using the Stmctural Relationship

The mixture model for the density (eq.2.1) provides an explicit representation for the

ratio of the specific volumes of H2O and D2O at pressure P and temperatures T and T',

83
VjPJ^ ^ f/.PXH)V/,PXH) ^ {fjj(PJMyjj(PXff)
V{PJ',D) ff,Pj'.D)V^{Pj',D) + [fj^Pj'mVi^PJ'^D) ^^^^

where V, and Vn are the specific volumes ofthe assumed two bonding types I and II in the

mixture-model picture, and f, and fn are the bonding fractions of the type I and II

components. According to the temperature scaling hypothesis the bonding fraction in H2O

equals the bondingfractionin DjO when T' = T + 6, where 6 = 7.2 .

Assuming that the stmctural properties of type I as a function of temperature, T

versus T', are the same for both isotopic molecules, leads to the following relationship

pfP,T'jy) = R,9/P,TJI) . (4.2)

The two densities are related by a constant factor R,,, where the ratio R^ refers to the shifted

temperature scale, T' = T + 6. Similarly, assuming that the same relationship with the same

constant factor holds for the type II bonding stmctures

Pjj{P,Pj)) = R^Pjj{P,Tjr) . (4.3)

The constant R^ should be close to but not equal to the ratio of the molecular weights,

20.023/18.010565 = 1.1117, because of slightly different hydrogen-bond distances and

atomic root-mean-square displacements within the same stmctural constraints [146].

Utilizing the relationships in eqs.4.1-4.3 a generalized equation for the liquid densities can

be written.

84
(4.4)
p{p,m
Rs, is called the scaled density ratio. The density ratio obtained when no temperature shift

is applied is the unsealed density ratio, \

9(P,Tj:)) _ ^
(4.5)

The density of liquid H2O [18] and DjO [17,117,133] from -30 to +50 C and the

corresponding unsealed density ratios Ry are presented in Table 4.1. The density of liquid

H2O [18]from-30 to +50 C, and the density obtained from Kell'sfittingequation [17], of

D2O at the shifted temperatures [t(H20) + 7.2] C along with the scaled density ratios R^. are

provided in Table 4.2. As mentioned before, the density of H2O is very likely known to

around five significant figures, that of D2O is known only to about four significant figures.

The ratio R^ of these densities for temperatures between -30 C and +30 C, as seen

from Table 4.2 on the H2O scale is 1.1060 0.0001. The small variation in the R^ values

is very probably caused by the slight D2O density uncertainty, particulariy in the lower

temperature range ofthe supercooled regime [133]. This constant ratio on the shifted

temperature scale is in contrast to the unsealed density ratio R^ ratio, shown in Table 4.1,

when no temperature shift is applied.

As seenfromTable 4.1, the R^ ratio increasesfrom1.0914 at -30 C to 1.1081 at +30

C, a variafion that is over two orders of magnitude larger than that of R^. Above about +30

C, the ratio Rj begins to deviate more from its low temperature value. This is caused, as

85
Table 4.1 Unsealed Densities of D,0

t(H20) C p(H20),t p(D20),t K


-30 0.98385 1.0738 1.0914
-25 0.98959 1.0845 1.0959
-20 0.99355 1.0918 1.0989
-15 0.99628 1.0970 1.1011
-10 0.99812 1.1006 1.1027
-5 0.99926 1.1031 1.1039
0 0.99984 1.1047 1.1049
+5 0.99996 1.1056 1.1057
+10 0.99970 1.1060 1.1063
+15 0.99910 1.1059 1.1069
+20 0.99820 1.1053 1.1073

+25 0.99704 1.1045 1.1077

+30 0.99565 1.1032 1.1081

+35 0.99403 1.1017 1.1083

+40 0.99222 1.1000 1.1086

+45 0.99021 1.0979 1.1088

+50 0.98804 1.0957 1.1090

86

tAtS'VK' - . ' . * 3 l l * W i
Table 4.2. Scaled Denshies of D,0

f = tOHjO) + 7.2 C
discussed in the previous chapters, by different thermal expansivities of the I and II

components, a change in the stmcture ofthe type II component to more varied bonding

forms, and eventually to a breakdown ofthe entire two-state mixture-model concept with

increasing temperature.

4.2. Scaling Prncpdure Applied to thp Density of T2O

In principle, the scaling idea should also work for the density of T2O. Goldblatt

[147] has reported reliable density data for liquid T2O from +5 to +54 C. Initially the

density of 99.30 mole% T2O, prepared and analyzed by Goldblatt [147], was determined by

the magnetic float method [148]. The density values for 100% TjO were then calculated

using Longworth's equation [149]. The maximum density is reported to be 1.21502 g cm'"^

near 13.4 C. Thus, for T2O, temperatures must be shifted by about 9.4 C relative to HjO.

4.2.1. Densities of H.O and T>,0


2^

The density of liquid H2O [18]from-15 to +50 C, the density of T2O at these same

temperatures are presented in Table 4.3 along with the unsealed density ratios R^, calculated

for each temperature listed. The densities of T2O at different temperatures in Table 4.3 are

calculated using the empirical function given by Kell [17]. The density of liquid H2O [18]

from -15 to +50 C, die calculated density of TjO at the shifted temperatures [t(H20) + 9.4]

C and the scaled ratio R^ of these densities for temperatures between -5.6 to +59.4 C are

given in Table 4.4.

88
Table 4.3 Unsealed Densities of T^O

t(H20) C P(H20),t p(T20),t


-15 0.99628 1.2041 1.2086
-10 0.99812 1.2080 1.2103
0.99926 1.2109 1.2118
0 0.99984 1.2129 1.2131
+5 0.99996 1.2142 1.2142
+10 0.99970 1.2149 1.2153
+15 0.99910 1.2150 1.2161
+20 0.99820 1.2146 1.2168
+25 0.99704 1.2138 1.2174
+30 0.99565 1.2125 1.2178
+35 0.99403 1.2109 1.2182
+40 0.99222 1.2090 1.2185
+45 0.99021 1.2069 1.2188

+50 0.98804 1.2044 1.2190

89
Table 4.4. Scaled Densities of T^O

t' = t(H20) + 9.4 C p(H20),t p(T20)>t'


5.6 0.99628 1.2106 1.2151
-0.6 0.99812 1.2127 1.2150
+4.4 0.99926 1.2141 1.2150
+9.4 0.99984 1.2148 1.2150
+14.4 0.99996 1.2150 1.2150
+19.4 0.99970 1.2147 1.2151
+24.4 0.99910 1.2139 1.2150
+29.4 0.99820 1.2127 1.2149
+34.4 0.99704 1.2111 1.2147
39.4 0.99565 1.2093 1.2146
44.4 0.99403 1.2072 1.2144
+49.4 0.99222 1.2048 1.2142
+54.4 0.99021 1.2021 1.2140

+59.4 0.98804 1.1993 1.2138

90

'^-<,'.*iXi^mt0mmHat
The scaled density ratio, R for temperatures between -15 and +20 C is 1.2150

0.0001. This scaled density ratio remains constant over a narrower range of temperature

when compared to the scaled density ratio of H2O and D2O. Above +30 C, the ratio tends

to gradually decrease. On the other hand, the unsealed ratio steadily increases from 1.2086

at-15 C to 1.2190 at+50 C.

4.3. Application to Other Properties

This idea would be relevant not only to densities, but also to other properties such

as the radial distribution fiinctions [150], and possibly also to the dynamics [151], or any

other property, such as viscosity, that depends solely on the stmcture, once any effect ofthe

mass difference, 18.010565 for H2O and 20.023 for D2O, has been accounted for.

The shear viscosity of H2O for twenty-six temperature points between -4 and -35 C

is the same as that of D2O at [t(H20) + 7] C (within a few tenths of a centipoise) [125].

The viscosities vary from 2 to 19 cp in this temperature range. However, the temperature

scaling ofthe more precise H2O and D2O viscosity data at temperatures higher than -4 C

gives an isotopic ratio of q(D20)/q(H20) of about 1.05. This value is close to what the

square-root-mass law would predict.

This phenomenon should also be usefijl in correlating some ofthe properties ofthe

isotopic forms of water with those of H2O, thus providing a check of their validity to the

much higher precision available for H2O data. Properties such as melting and boiling points

of H2O and D2O do not behave similariy because they have a thermodynamic rather than a

stmctural basis.

91
This very precise temperature-scaling law provides more credence for the mixture-

model approach since no parameter variations are involved here.

92

!-*:*- rjrjiissaa
CHAPTER V

THE ISOTHERMAL COMPRESSIBILITY MINIMUM NEAR 50 C

AND OTHER ANOMALOUS PROPERTIES

5.1. Qualitative Description ofthe Isothermal Compressibility

Normal liquids have compressibilities at 1 atm pressure of approximately 0.1 kbar'

at room temperature. The compressibility of ice at 0 C is 0.013 and the compressibility of

water at 0 C is 0.051. Thus, water has roughly half the compressibility of normal liquids,

though it is considerably more compressible than ice. Water also differs from normal liquids

in the temperature dependence of its compressibility. Typically the compressibility of a

normal Uquid nearly doubles in going from 0 C to 100 C. In the case of water, the

compressibility decreases by about 13% in going from 0 to 45 C, . After passing through

a minimum at 46.5 C the compressibility increases, but at 100 C it has still not quite

returned to the value at 0 C.

Rontgen [21] applied his mixture-model concept to provide an explanation for the

observed minimum of compressibility. According to Rontgen's hypothesis, if pressure is

applied to water at a fixed temperature above its freezing point, the ice-like (ice-I-type)

molecules tend to transform into the more dense second (ice-II-type) type of molecules. The

observed decrease in the volume of water caused by pressure is because ofthe simultaneous

occurrence of compression and contraction (due to the above mentioned transformation).

At low temperatures, the effect of the actual compression is possibly smaller than the

transformation into dense molecules. This is because there are more ice-like (or ice-I-type)

93

T . * = V i * ? * * * - --. ^ -*v -^^


molecules at lower temperatures. Therefore, there exists a minimum of compressibility in

the vicinity of 50 C.

Quantifying Rontgen's hypothesis would lead to the expression given by eq.2.1 given

for the pressure/temperature dependent specific volume of liquid water. As already

mentioned, this expression is the basis ofthe modernized mixture model proposed here and

the terms involved in this expression are explained in Chapter II. The isothermal

compressibility factor K(T) is represented as,

K(7) = -\IV,^{dV,JdP), . (5.1)

Using eq.2.1, the isothermal compressibility factor may be written as a sum of three terms:

K{T) = i/K,/r,p) ((/; K/j)F/r,p)


* (f,i hfT)V,fT,P)) (5 2)
* (df, IdP) [V,fT,P) - V/(T,P)])

All the three terms are positive. Thefirsttwo terms are positive since fi(T,P), K,(T)

and KII(T) are positive and the third term because dfi/d? and (Vn - Vj) are negative. Looking

at eq.5.2 it is difficuh to assess how a minimum in the compressibility factor could come

about when all the three contributing terms are all positive. The 'proportionality effect' that

caused the density maximum in liquid water is also responsible for the minimum in the

isothermal compressibility. This may be explained as follows:

At very low temperatures, the third term in the above expression is relatively larger

than the sum ofthe first two terms. This term, as already emphasized in ref [41], quickly

goes to zero as the temperature is raised, partly because of the difference between the

94

rm^^^a^jwa
volumes of the two components, (VQ - V,). But this is mainly because the pressure

dependence of fj strongly decreases with increasing temperature. In fact, at temperatures

higher than about 350 K, the third term almost diminishes to zero as the order parameter, f,,

becomes pressure independent at such temperatures [152]. The first term steadily, but

slowly, decreases as the temperature rises due to the decrease in f,(T,P) itself

Simultaneously the second term increases in magnitude a little too fast than the decrease in

the first term. This leads to a weak increase of their sum with rising temperature. However,

this weak increase does not compensate for the large decline in the magnitude ofthe third

term. Thus, anomalously, the isothermal compressibility of liquid water decreases at low

temperatures because of the "proportionality" of the dominant third term decreases in

magnitude with increasing temperature. Calculated values for the three terms will be given

in the next section of this chapter.

5.2 Quantitative Description ofthe Isothermal Compressibilitv

It is evident from the description given in the previous section that it is the decreasing

magnitudes of fi(T,P) and d^i/dP with increasing temperatures that is responsible for the

thermal minimum of K(T) in liquid water. Hence, accurate order parameters must be

available, both at 1 atm and elevated pressures, in order to provide a convincing quantitative

discussion of K(T) given in eq. 5.3. From the values of fi(T,P), the needed derivatives af/aP

can be determined for a range of temperatures. In the absence of an ab intio calculations,

these derivatives in principle can be obtained from the temperature and pressure dependent

activation energies for dynamical processes, such as, viscosity, spin-lattice and dielectric

95

I B U B B i f W ! i| 4W'-<lL'-'-t;j.'JIAJi
relaxation [41]. The ^'0 spin-lattice relaxation times measured by Lang and Ludemann

[153] are able to provide cmde values of af/ap, which were used for the analysis in ref [41].

The data of ref [153] are reported to no more than 3 significant figures. Considering the

inherently much higher accuracy of density measurements (up to six to seven decimal

places), it would be much better to obtain the derivatives directly from the measured density

of water as a fijnction of T and P. However, searching for literature about the pressure

effects on the density of liquid water, little high precision information is available in the

important supercooled regime. Therefore, experimental data used here will be over a

narrower range of temperature and of a lower precision, compared with the data employed

in the density analysis discussed in Chapters II and III.

5.2.1. Experimental Data

The change of density of liquid water under pressure has been calculated fairiy

accurately by Kell [154] and by Walker [155], from the analysis ofthe speed of sound. At

frequencies, where dispersion is not important, the speed of sound u is related to the

isentropic change of density with pressure, K^ (isentropic compressibility or adiabatic

compressibility). Fitting u"^ as a polynomial in temperature and pressure, then integrating

with respect to pressure, the density of water under pressure can be calculated. The

isothermal compressibility was calculated by Del Grosso from the values of K densities,

thermal expansitivities and heat capacities at various temperatures [154,156,157]. The

calculated data are represented by a seven-parameter rational fijnction by Kell [154] and he

uses this function to recalculate the values of K(T) at different temperatures. Kell reports

96
m- I m'f'm

K(T) values to six significant figures from 0 to +50 C, at 1 atm pressure. Values of these

factors for temperatures below 0 C down to -30 C obtained from extrapolation of the

empirical ftinction have also been listed by Kell [18] to four significant figures. Above +50

C, they are reported to five significant figures. Walker has reported two sets of isothermal

compressibility data and has listed them to five significant figures from 0 to +100 C.

Comparing the data of Kell and Walker between 0 and +50 C, it was found that the

data often disagree in the third significant figure. The somewhat more recent data of Kell

was utilized for this analysis, but only four significant figures were considered ofthe Kell

data, in the region of interest here, which is 0 to +50 C. Above +50 C, only three

significant figures were considered since the agreement between Kell's and Walker's data

occurs only up to two significant figures. In fact, above +70 C, the Walker values are

consistently lower than Kell's. Hence, only Kell's values up to +70 C were considered.

Values of K(T) for selected temperatures from -30 to +70 C are listed in Table 5.1.

It is clear that the values start decreasing from 80.8 xlQ-^bar^ at -30 C to 44.15 xlQ-^bar'

at +45 C. It remains at 44.15 xlQ-^bar"^ from +45 through +48 C (a very shallow minimum

is exhibited), and from thereon it increases steadily with rising temperature and reaches a

value of 49.02 xlO-^bar^ at +100 C.

97
Table 5.1 Observed and Calculated Isothermal Compressibility Factors as a Function of
Temperature.

t,C K {T)\ obs K (T)', calc


(-30) 80.8 73.4
(-25) 70.9 67.6
(-20) 64.3 62.9
(-15) 59.4 58.9
(-10) 55.8 55.7
-7 54.08 54.04
-5 53.06 53.05
0 50.89 50.91
5 49.17 49.18
10 47.81 47.80
15 46.73 46.72
20 45.89 45.88

25 45.25 45.24
30 44.77 44.78

35 44.44 44.46

40 44.24 44.26

"Compressibility factors are in units lO-^bar'V

Temperatures outside thefittingrange -7 ^ t ^ +50 C are in parentheses.

98
Table 5.1 Continued

t, c K (T)*, obs K {jy, calc


42 44.19 44.20
44 44.16 44.17
45 44.15 44.16
46 44.15 44.15
47 44.15 44.14
48 44.15 44.14
49 44.16 44.14
50 44.17 44.14

(55) 44.3 44.2


(60) 44.5 44.3

(65) 44.8 44.5

(70) 45.2 44.7

"Compressibility factors are in units 10 bar".

Temperatures outside thefittingrange -7 <. t ^+50 C are in parentheses.

Values in bold indicate where the minimum occurs for the observed and calculated
compressibility factors.

99
5.2.2. The Fitting Procedure

In thefittingprocedure employed here, only experimental data from -7 to +50 C was

utilized. Isothermal compressibility factors from the supercooled regime were not included

because experimental data in this region was obtained by extrapolation and possible

inaccuracies might exist in the values thus obtained [18]. Values up to +50 C were chosen

to show how the compressibility minimum in the vicinity of+50 C arises. K(T) values

above +50 C were not be included in the analysis because at higher temperatures the

mixture model used must certainly break down because of varietal and disordered bonding

coming into play in the stmcture of liquid water. Therefore, only 58 data points from -7 to

+50 C were considered for this purpose.

K(T) was related to the six unknown functions -Vi(P,T), Vn(P,T), fi(T,P), KI(T), Kn(T)

and af/aP- for each temperature T and pressure P. Expanding these six parameters

analytically as fiinctions of T and P would involve a large number of unknown quantities.

Since the experimental data of Kell presented in Table 5.1 are for 1 atm pressure, only the

temperature dependence ofthe six functions is required. Three ofthe six fijnctions -Vi(P,T),

Vn(P,T), fj(T,P)- are known from the liquid H2O denshy analysis discussed in Chapter II.

Therefore, only the three remaining fiinctions -KJ(T), Kn(T) and af/aP- are to be determined.

The compressibility factor, KJ, corresponds to the pure type-I component of liquid

water and K^ corresponds to the pure type-II component. The magnitudes of KJ and K^

might be somewhat larger than the compressibilities [10,158] of ice-Ih (12-21 xlO"^ bar"' at

-7 C ) and ice-II (unknown) or ice-VI (4-6 xlQ-^ bar'' near 0 C). These compressibilities

should behave like those of normal liquids and therefore would be expected to increase with

100
temperature. The temperature dependence of K(T) for most pure substances is fairiy weak,

hence KI(T) and Kn(T) for liquid H2O should be fairly weak too. These two fiinctions
can
be expanded linearly as aftinctionof temperature.

^/7) = K/T;) [1 . y/r-z;)] (5.3)

'^//^ = ^/X^o) [1 ^ yj/iT-T^] (5.4)

Eqs.5.3 and 5.4 introduce four new parameters K,(TO), KU(TO), YI and YD where the y's are

poshive. TO is essentially the same parameter as in the previous chapters and the value is ~

225 K.

The next task is tofinda suitable analyticalftinctionfor the derivative afj/aP. Based

on the trend, as seen in Fig. 5 {af/aP is represented by the ratio AH(T,P)/AH(To,P)) of ref

[41], the magnitude ofthe derivatives, af/aP, are large at low temperatures, it decreases

{df,/dP)^ = -AexpirKT-T^y (5.5)

as the temperature increases andfinallygoes to zero near 350 K. Since the change of af/aP

with pressure is not linear due to 'cooperativity effects,' the function may be represented by

a given in eq.5.5. qs.5.3 through 5.5 introduces six fitting parameters for the 58 data

points employed in this analysis.

5.2.2.1. Preliminarv Fits

A least-squaresfittingprocedure (see the Appendix) was adopted to produce a best

fit of all the experimental data points. Varying all six parameters simultaneously resulted

101
in ambiguous output values for these parameters since apparently several minima exist in the

determination of x' which gives good fits for the data. Some such output values are: (i)

negative values of one ofthe y's, compensated by too large a positive value for the other, (ii)

negative values for one ofthe K'S, with an unrealistically large positive value for the other

K. In order to obtain realistic output parameters, some ofthe initial parameters were fixed

at certain values and the rest were varied. Fixing certain parameters did not affect the x"

values and the experimentally observed isothermal compressiblity minimum almost always

occurred in the recalculated data. But the temperature at which the minimum occurred

always depended on the output parameter values. It was found that, almost always, the

isothermal compressibility minimum occurred a few degrees above the experimental

temperature of 46.5 C.

5.2.2.2. Least-Squares Fit (5A)

This fit of Kell's K(T) data in the temperature range -7 to +50 C employs only four

adjustable parameters. The volume parameters given in eq 5.2 -Vi(P,T), VQ(P,T), fi(T,P)-

are calculated from the fit (2.4) parameters ofthe liquid H2O density analysis discussed in

Chapter II; these values at different temperatures in the specified range are given in Table

5.2. A number of preliminary fits showed that the values ofthe two parameters, Kn(To) and

Yi, were close to 1.50 x 10 bar' and 3.00 x 10'^K-' respectively Hence, for this fit, the

parameters Kn(To) and YI were fixed at these values. The output pressure parameters

obtained from this analysis are presented in Table 5.3 along with the volume parameters

102
Table 5.2. The Volume Quantities inEq.5.2 as a Function of Temperature*

t, C Viic' Vi^ Vn^ fi^


-30 1.0164 1.0963 0.8670 0.6518
-20 1.0065 1.1012 0.8779 0.5757
-10 1.0019 1.1062 0.8889 0.5199
-5 1.0007 1.1087 0.8944 0.4963
0 1.00016 1.11118 0.89987 0.47460
5 1.00004 1.11367 0.90535 0.45451
10 1.00030 1.11616 0.91083 0.43573
15 1.00090 1.11865 0.91632 0.41806
20 1.00180 1.12114 0.92180 0.40134
25 1.00296 1.12362 0.92728 0.38547
30 1.00437 1.12611 0.93276 0.37037
35 1.00600 1.12860 0.93824 0.35597
40 1.00785 1.13109 0.94372 0.34223

45 1.00988 1.13357 0.94921 0.32911

50 1.01211 1.13606 0.95469 0.31659

55 1.0145 1.1385 0.9602 0.3046

60 1.0171 1.1410 0.9657 0.2933

65 1.0198 1.1435 0.9711 0.2825

70 1.0227 1.1460 0.9766 0.2723

" The units are as given in Chapter II.

^ Vuq= experimental molar volume ofthe liquid.

' Vi and Vn calculated from the volume parameters given in Table 5.3,

^f, calculated from eq 2.1 using V^^, Vj and Vn

103

."^ V / * : - r-,:
Table 5.3. Volume Parameters from the Density Analysis and Pressure Parameters from the
Least-Squares Fit (5.1)

Volume Parameters' Density Analysis Fit (2.4)


V,(T) 1.08739
Vn(T) 0.84745
"i 4.5756 -4*"
n 1.2937 -3
T 225.334
Pressure Parameters'^ Least-Squares Fit (5.1)
Ki(To) 3.7721 -5
KnCV 1.50 -5

Y/ 3.00 -7

Yn 1.3360 -7
A 3.0054 E-4
k 3.2626 -2

The units are as given in Chapter II.

"4.5756 -4 should read as 4.5756 x 10"

" The units are as given in the text.

** Parameter fixed.

104
parametersfromdensity analysisfit(2.4) which were used to calculate to V,^(T,P), V,(P,T),

Vii(P,T) and fi(T,P). The recalculated isothermal compressibility factors in the temperature

range -30 to +70 C from this least-squares fit is provided in Table 5.1 along with the Kell

data. The three contributing terms and the derivative af/aP in eq.5.2 were calculated from

the output parameters obtained from this fit and are displayed in Table 5.4.

5.2.3. Discussion

The reason for the occurrence of the minimum at a temperature which is a few

degrees above the temperature at which the experimental minimum occurs could be partly

due to fits obtained by the use of too few parameters. In the density analysis, to fit the

higher temperature data (above +40 C) the quadratic thermal expansion term, p(T-To)^, was

included. This term was omitted in the current analysis. Poor fits are generated not only by

the use of too few parameters leading to the wrong curvature ofthe calculated K(T) values,

but also because ofthe temperatures near the point where the minimum occurs being at the

tail end ofthe experimental K(T) list in the density analysis.

Examining the calculated K(T) values at the two extreme ends ofthe temperature

range, one finds that these calculated values are slightly lower than the corresponding

experimental values. This indicates that morefittingparameters or a different form of eq.5.5

may be necessary to give a higher curvature. The shallowness of the curve could be another

reason for the difficulty experienced in reproducing the temperature ofthe minimum. Also,

as mentioned before, greater uncertainties occur not only in the mixture model at the higher

temperatures, but also in the experimental data.

105

iiiijMi.<!!ifijijmmg
^fr iMon^u^tmim
Table 5.4. The Three Contributing Terms and af/aP in Eq.5.2 as a Function of
Temperature"

t,C af/ap Term 1 Term 2 Term 3


-30 168.05 30.27 5.16 37.91
-20 121.27 29.02 6.93 26.90
-10 87.51 28.17 8.54 18.98
-5 74.34 27.80 9.33 15.92
0 63.149 27.454 10.111 13.342
-53.643 27.113 10.892 11.175

u 10

15
45.569

38.710
26.773

26.430
11.676

12.463
9.354

7.825
20 32.883 26.080 13.254 6.543
25 27.934 25.723 14.050 5.468

30 23.729 25.359 14.850 4.568

35 20.157 24.986 15.655 3.814

40 17.123 24.606 16.465 3.183

45 14.546 24.221 17.278 2.656

50 12.356 23.833 18.094 2.214

55 10.50 23.44 18.91 1.85

60 8.92 23.05 19.73 1.54

65 -7.57 22.67 20.55 1.28

70 -6.43 22.29 21.36 1.07

" af/ap and Terms 1, 2 and 3 have units of bar* x 10"

106

" f i I I ' ! . > >


Although, the compressibility is not exactly reproduced using the least-squares fit

parameters, the differences between the calculated and the observed values of K(T) in the

fitting range are mainly one unit in the last reported decimal place. As already mentioned,

a noticeable negative error occurs at the two ends of the temperature scale used. It is

estimated that the error at both -7 C and +50 C is around -0.07 %, while, in the middle of

the temperature range, errors fluctuate near 0.02%. Table 5.1 also includes experimental

and calculated K(T) values outside the narrower fitting range employed for the analysis

These values show to what extend deviations occur when using fitting parameters from the

-7 to +50 C temperature range.

5.3. A Note About the Heat Capacity. Dynamic Properties and Viscosity

More experimental work on the heat capacity and intermolecular vibrational

spectmm of liquid water in the supercooled region at normal and elevated temperatures is

needed if the mixture model described here is to be used to explain the abnormal behavior.

The fact that certain librational frequencies in liquid water strongly increase with

increasing temperature [159-161] can easily account for the anomalously high heat capacity

[41]. Also, since fi(T,P) decreases with pressure as more ofthe weakly bonded, dense

component is produced, these frequencies would also be expected to decrease with

increasing pressure, increasing the heat capacity contributions from these modes. Increasing

the pressure should additionally moderate the temperature dependence ofthe librations, so,

even though the heat capacity tends to be larger as the pressure increases, the anomaly ofthe

107

jMMHaaHSKij
temperature dependence of the heat capacity should become less prominent at such

pressures.

The intermolecular potential surfaces in liquid water, cooperatively flatten out and

soften as the temperature is raised. Simultaneously, the local bonding in liquid water

changesfromthe open type-I to the dense type-II bonding. There should exist a relationship

between this bonding and the dynamic properties of water, including the viscosity. The

softening ofthe intermolecular potential surfaces [41] is believed to be responsible for the

decreased activation energies [153,161-165] for dynamical processes in water as the

temperature or pressure increases. As already mentioned, another evidence concerns the

observed lowering ofthe librationalfrequencieswith increasing temperatures [159-161 J.

The pressure effects responsible for increasing the dynamics in liquid water also provide an

explanation [10,19] for the decrease in the viscosity when pressure is raised at temperatures

less than about 20 C.

108

gi!^'ig<inawPiiiKst^-.. wii^ ^mf^*m^mmamtm^


.t..^-^iteAa ism
r
CHAPTER VI

INDEPENDENT SUPPORT FOR THE MIXTURE MODEL

A stmcture for liquid water based on ice-like arrangements of water molecules does

require contributions from dense ice stmctures, as pointed out from our density analysis and

isothermal compressibility considerations. Such contributions are expected for a substance

showing extensive polymorphism at relatively modest pressures. Because of the labile

character ofthe liquid, due to its relatively large thermal motions and the rapid changes in

molecular configuration taking place, one would expect a number of molecular arrangements

to be present in the Uquid. This is possible because the corresponding solid shows, by its

rich polymorphism, that a variety of molecular arrangements with roughly comparable

energies can be constmcted. In the following sections, some evidence is provided by

stmcturally controlled properties of water for the occurrence of two types of bonding forms

ofthe types described in earlier chapters.

6.1. Radial Distribution Function

X-ray diffraction provides geometrical information on water stmcture, in the form

ofthe average distribution of oxygen-oxygen distances r occurring in the liquid. The radial

distribufion function (r.d.f) determined by Morgan and Warren [26] for liquid water at 1.5

C is shown as the lowermost curve of Fig. 9 in ref [15]. For comparison with water in Fig.

9 are theoretical r.d.f's for known and hypothetical forms of ice, calculated from the discrete

109

MMMI
distributions of mean oxygen-oxygen distances in the ice crystals by applying a Gaussian

smoothing to represent the effects of thermal motion.

The primary feature ofthe water r.d.f is the presence of about four oxygen neighbors

at distances near r = 2.8 A. This indicates a basically tetrahedral coordination for the liquid

water molecules, similar to the coordination in most of the ice forms. Comparing the

observed r.d.f for water whh calculated r.d.f's based on the ice I stmcture, Morgan and

Warren [26] noted that,

1. The nearest-neighbor peak is shifted from 2.76 A in ice-I to about 2.9 A in water,

2. The next-nearest-neighbor peak near 4.5-5.0 A is broader and lower in water than in ice-I,

3. In the interval between 3.0 and 4.0 A in the water r.d.f there is considerable excess of

oxygen-oxygen distances, corresponding to a peak centered at about 3.6 A.

Comparing the r.d.f of liquid water whh those for the dense forms of ice, Kamb [ 15]

noted that all of the above features found in water: the increase in H-bond lengths, the

variable distances of next-nearest neighbors, and the presence of numerous non-nearest

neighbors in the interval 3.0-4.0 A, correlate in a general way with what is found in the

denser ice phases.

6.2. Kamb's work

In order to prove further the point that the r.d.f for water can be accounted for by the

presence in the liquid of ice-like molecular arrangements, Kamb [15] compared the r.d.f for

liquid water with theoretical r.d.f for a stmctural mixture consisting of 50 percent ice-I, 33

percent ice-II, and 17 percent ice-in. The theoretical curve, containing different proportions

110
of ice -I, -II and -III, (see Fig.lO in ref [15]) matches fairly well, probably to within

experimental uncertainty, the observed r.d.f for liquid water at +4 "C.

In fact, from the density analysis done on liquid H2O, it was shown (see Table 2.10)

that the fraction fj of ice-I-type bonding, calculated from the fit (2.1) density analysis, is

-0.69. At +4 C this fraction drops to -0.49, this agrees very well with the proportion of

ice-I chosen by Kamb [15] to form the stmctural mixture of liquid water at the same

temperature.

6.3. Intemal Vibrational Frequencies

As mentioned in section 1.3.1.1. of Chapter I, Walrafen [57] has conducted a large

number of Raman studies of liquid water and his investigations strongly favor the mixture

model theory in general. The Raman spectmm of liquid water can be approximately

resolved into two components [57]; one component has been termed 'hydrogen-bonded' and

the other as 'non-hydrogen-bonded.' This has been confirmed by others [166-168].

Moreover, Walrafen [59] studied the Raman spectral data of Frank and Linder [169] which

stretched over the temperature range 16 to 100 "C and made some significant observations

in favor ofthe mixture model. Walrafen [59] found that the data of Frank and Linder [169]

could be decomposed into two Gaussian components. He found that the frequency ofthe

non hydrogen-bonded component remains relatively constant (2644 - 2649 cm "') over the

entire temperature range. This suggests the existence of a temperature-dependent

nondirectional intermolecular perturbation of the OH stretching vibration. Another

significant observation was that the frequency of the hydrogen-bonded component

111
increases from 2516 to 2618 cm"' over this temperature range. This shift is in the direction

of hydrogen bond weakening. Such a weakening might occur in several ways: an increase

in the O-H-0 bond length, consecutive hydrogen-bond breaking or a bending ofthe bond.

An increase in the O-H-0 bond length at constant density must be accompanied by an

increase in the coordination number. However, it is known [49] that from 0 to 200 "C the

coordination number in water remains relatively constant. Thus, it would appear that one

ofthe last two explanations is more likely. Although Walrafen favors bond breaking, our

mixture model theory combined with the extensive stmctural data on the ice polymorphs

from Kamb's work [15] most definitely favors bond bending. The average hydrogen bond

length (Table 1.2), which is an inverse measure ofthe bond strength, increases progressively

from 2.75 A in ice-Ih to 2.81 A in ice-VI. Bond length changes thus work against the effect

of bond bending in achieving denser molecular packing in these high-pressure forms [15].

From the low-temperature Raman studies [168], it is clear that the lower frequency

component in supercooled liquid tends toward the ice-Ih or amorphous ice-I Raman

spectmm as the temperature is reduced. These aspects also seem to support the mixture

model proposed here. However, it remains to decipher the entire temperature-dependent

infrared and Raman spectmm of liquid water at low temperatures in terms ofthe ice-Ih/ice-II

model. This chore is made difficult by the probable presence of temperature-weighted ice-I-

type and ice-II-type infra- and intermolecular modes, including phonon modes, whose

temperature-dependentfrequenciesare coupled together [168], with the coupling parameters

themselves probably being temperature dependent.

112
^^i^-Slngujar Behavior .nearj;^

At the beginning of our analysis, a variable power law parameter n was incorporated

into the fitting ftinction fn(P,T) of eq.2.6 so that this ftinction would approach zero as (T -

To)". However, preliminary fits immediately showed that n was extremely close to unity, so,

subsequently, n was set to one. Such a power law dependence shows that the ftmction

f(P,T) is not singular; i.e., none of its derivatives diverge as T approaches T,. Other

published work [88,170] indicated otherwise. However, ref 121, finding no strong increase

in correlation length in the vicinity of TQ, also concludes that there is no singular behavior

near this temperature. It should be remarked here that the cooperative melting phenomenon

discussed in ref 41 does not demand a singular ftinction for the fraction of 'rotating'

molecules. The important feature for the water problem is the effect that these rotating

molecules have on softening andflatteningthe intermolecular potential ftinctions, somewhat

reminiscent of 'phonon mode softening' in soHd-state physics [171].

6.5. Stmctural Relationships

Comparing the illustrations of ice-Ih and ice-II (Figures 1.1 and 1.2), one sees that

in ice-II every other hexagonal channel in the ice-Ih stmcture has been filled in with a

collapsed stmcture having roughly twice the density ofthe open hexagonal stmcture. In ice-

II the walls ofthe hexagonal tunnels are composed ofthe collapsed stmcture. This is the

origin ofthe 22% loss of molar volume in the ice-Ih to ice-II transformation. In the denser

forms of ice, probably some stabilization energy is gained because of the near favorable

O-O van der Waals distance of-3.3 A in these stmctures. In the liquid state, compared with

113
the ice-II crystal, the free energy of component II is probably lowered, since the H atoms

making up the hydrogen bonds are very likely more disordered as far as to which ofthe tuo

O atoms they belong. The entropy of proton disorder [172] in the ice-Ih crystal provides an

important contribution to the polymorphs stability.

An interesting description of the stmctural relationship between nine of the ice

polymorphs, including ice-Ih, -II, -III, -V, and -VI, has recently been presented by Dmitriev

et al. [39]. They have described the stmctures of nine phases of ice in terms of 'ordering and

displacive' mechanisms from a common parent, denoted as IQ, possessing a disordered body

centered cubic stmcture. They showed that IQ was the maximal substmcture common to all

ice stmctures provided different concentrations of water molecules are used in making the

nine different ice stmctures. One can, in fact, picture a fluctuational network of ice-I-type

bonding interlaced with ice-II-type bonding (plus ice-III, -V, or -VI types at elevated

temperatures or near certain interfaces).

6.6. Isochoric Temperature Differential ofthe X-Ray Stmcture Factor

Recently, Bosio, Chen and Teixeira [122,123] have studied the temperature

dependence ofthe x-ray stmctural parameters from isochoric temperature differentials. X-

ray stmcture factors for Hquid D2O were measured at temperatures 40.0, 23.5, 15.5, 11.2,

7.1, 0.2 and -11.0 C [123]. The x-ray stmcture factors reflect mostly the molecular center

ofthe oxygen local positional (0-0) correlations with a small admixture of 0-D correlations.

From these data, isochoric temperature differentials were constmcted for pairs of

temperatures, AT = 40.0(T,)-(-11.0)(T2) =51.0, AT = 23.5 - 0.2 = 23.3, and AT = 15.5-7.1

114
= 8.4 C around the maximum density point of 11,2 C. T, and T, being the two

temperatures on each side ofthe density maximum for which the density is the same.

The derived quantity ofthe x-ray stmcture factor is the pair correlation ftinction g(r).

Using the differential quantity, Ag (r, AT), the temperature variation ofthe number of

neighbors AN(R) up to a distance of Rfromthe central molecule can be obtained accurately.

The resuhs of their study can be summarized as follows:

1. The first coordination shell, the area around the position ofthe first neighbor at 2.8 A,

is less strongly affected by the temperature variation than farther neighbor shells.

2. They noticed an important rearrangement of the molecular centers occurring at the

second-neighbor shell as the temperature is varied. A plot of Ag (r, AT) for the three

measured pairs of temperatures versus r shows the following features: The peak between

3.3-3.4 A grows in with increasing temperature differential. As already mentioned, this

peak does not occur at all in ice-I-type stmctures but is a prominent feature in the dense

forms of ice. Simuhaneously, there occurs a decrease in the intensity ofthe peak at 4.5

A. This peak corresponds to the non-nearest-neighbor O-O distance occurring in ice-I-

type stmctures but this does not occur at all in the dense ice forms.

The above resuhs seem to indicate that the local (0-0) positional correlation and the

0 - 0 - 0 angular correlations are at the origin ofthe density fluctuations and also of other

thermal anomalies of Hquid water. Also, they provide strong support for the existence in

Hquid water of a dynamic temperature-dependent mixture ofthe bonding types similar to that

occurring in the low and dense forms of ice. This is in exceHent agreement with the mixture

model theory proposed here.

115
In fact. It should be possible to combine the ice-Ih and the dense ice radial

distribution ftinctions in the proportions (f, and fn) that we have determined and roughly

reproduce, as did Kamb [15], the temperature-dependent radial distributionftinctionfor the

Hquid. The differential x-ray scattering curves of Bosio et al. [123] should also be possible

to reproduce by this technique. This stmctural analysis would constitute necessary and

important extensions ofthe mixture model concept for liquid water and this subject is being

investigated by Urquidi, Singh and Robinson, as part of on going research in our laboratory.

6.7. Computational Studies

It is evident [173,174] that until recently no computational model of water has been

able to provide a realistic explanation of the density maximum in water near 4 C.

Computational studies by Cho, Singh and Robinson [175] in our laboratory have provided

an explanation for this anomaly by ahering the water-water interaction potentials containing

an O-O Lennard-Jones contribution. Current water-water interaction models are mainly

concerned with the formation and dismption of the tetrahedral stmcture at the nearest-

neighbor level and the role played by the second-neighbor interactions are ignored. Cho et

al., have modified the Lennard Jones term in the water-water potential in such a manner that

when this potenfial is optimized, it wiH provide a second neighbor O-O minimum near 3.35

A, while keeping intact, under appropriate thermodynamic conditions, the open tetrahedral

stmcture. This is in accord whh the current mixture model for liquid water: the bonding

variations occurring in liquid water which are responsible for the anomalous properties are

116
not at the nearest-neighbor level, but instead occur in the outlying non-hydrogen bonded

second-neighbor 0 - 0 stmcture.

An exactly solvable one-dimensional Takahashi fluid model [176] was employed to

test the above proposal. A double-well potential was considered, the well farther from the

hard core being the deeper ofthe two. As would be expected, at zero pressure and T = 0 K,

the system exists in a crystalline state whh all the particles in the outer well. When the

temperature is increased at a fixed value of low pressure, the particles tend to 'boil' into the

inner well, increasing the overall density ofthe system. The results of this study [175]

confirm that a density maximum can be obtained for a one-dimensional water model at

constant pressure. On increasing the pressure in the one-dimensional model, the density

maximum becomes broader and shifts to lower temperatures and finally disappears at

sufficiently high pressures. This is the exact behavior of real liquid water under the influence

of pressure [16].

Results of ongoing research in our laboratory indicate that similar results can be

obtained for a three-dimensional double well potential model for water. In the three-

dimensional model, the outer potential minimum would correspond to the second-neighbor

at - 4.5 A distance, which occurs in ice-I-type stmctures. The inner minimum corresponds

to the second-neighbor distance at ~ 3.3 A occurring in the dense forms of ice and caused

by a bending ofthe hydrogen bonds [15].

This study [175] clearly emphasizes the need to insert empirically the appropriate

double-well feature into the water-water potential in order to explain the density maximum

in Hquid water. In the absence of such a modification, no density maximum can occur. The

117

iMMMIlfeHMBaHMMI
absence of a realistic density maximum has indeed been discovered to be the case for all the

popular computational models [173,174].

118
CHAPTER VII

LASER STUDIES OF THE PROPERTIES OF LIQUID WATER

PERTURBED BY SURFACES

It has been thought that a more stmctured form of water in confined volumes exists

and it has been suggested by various workers that the extent of this stmcture may range

from 10 A up to nearly I mm, with the stmctures varying from slightly ordered to ice-like

[93]. Because of these stmctural changes, certain ultrafast ionic reactions can be slowed by

orders of magnitude near an interface. Fluorescence probes whose characteristics depend

on these ionic reactions [177] should be able to distinguish such variations in the stmcture

of water. The probes l-naphthol and 2-naphthol function as proton-transfer probes and 8-

phenylamino-1-naphthalene-sulfonate (ANS)fijnctionsas an electron transfer probe, because

of a photon-induced charge-transfer-to-solvent (CTTS) process. In a stiff water

environment, where the orientational freedom of the solvent is suppressed, the charge

transfer rate is drastically curtailed.

In pure water, the charge transfer rate ofthe fast probe l-naphthol is 3x10' s , while

in ANS it is 5x10^s'\ and in 2-naphthol h is 2x10* s\ Competing with the fast charge

transfer in the exched states of these probes are the generally slower processes of ordinary

radiative and non-radiative decay. These decay times have little or no solvent or temperature

dependence. Because ofthe relative slowness and the need for thermal activation ofthe

ANS and 2-naphthol probes, only l-naphthol was used in the experiments to be described

119
here. The rapid charge transfer rate for this probe is attributed to the ability of water to

reorient freely and quickly into a proton-accepting water stmcture [177].

In the studies presented in this work, the total decay rate is measured. The rate is

directly related to the degree of stmcturing ofthe water present in the sample, and will be

seen to range between the bulk water value, where the charge transfer process is maximum,

to a value close to that in pure alcohol, where the process is completely blocked. Typically,

multi-exponentials are observed, whose lifetimes stretch between these two limits, but with

a clear separation of short- and long-lived components. All lifetimes reported here were

recorded at room temperature. In the present work, the decays are analyzed as double

exponentials, with the initial (t=0) intenshies providing the relative weights of the fast and

slow components.

7.1. Experimental Section

Thefluorescenceprobe l-naphthol (99.9%) was purchasedfromKodak. Water was

distiUed, deionized and passed through a NANOpure three-cartridge system. The measured

resistivity ofthefilteredwater is above 18.0 MQ/cm. Concentration of l-naphthol solutions

prepared in water was approximately 1.0 x 10-^ M.

Absorption and fluorescence spectra were measured on a Shimadzu UV-265

spectrophotometer and a SLM 4800Cfluorescencespectrophotometer. The emission bands

at 340 and 460 nm are attributed to the exched states ofthe neutral l-naphthol and the 1-

naphthol anion, respectively.

120
Fluorescence decays were measured with a time-correlated single-photon counting

(TC SPC) system. An Nd:YAG laser (Coherent Antares 76-S), a CW rhodamine 590 dye

laser and a second harmonic generator (SHG) mainly constitute this system. The Nd: YAG

laser produces a train of 70 ps pulses at a repetition rate of approximately 76 MHz with an

average output power of 2W at 532 nm. The average output power ofthe laser pulses from

the dye laser is about 180 mW at 610 nm. These pulses are frequency-doubled with a 50 mm

temperature tuned ADA crystal. The resulting 305 nm pulse with a 15 ps width is used to

exche the samples. The residual 610 nm radiation is removed by a UV pass filter.

The emissionfromthe samples was then selected by an appropriate interference filter

and collected by a fast response microchannel plate photomultiplier (Hamamatsu). The

output ofthe photomuhiplier was used as a start signal for the time-to-ampHtude converter

(TAC, ORTEC 457). The stop signal for the TAC is obtained from a photodiode which

detects the signals reflectedfroma UV pass fiher. The output pulses of different amplitudes

from the TAC were then scaled and accumulated by a multichannel analyzer (MCA, ORTEC

7010). The final spectral data were then stored in a PC for future analysis.

The instmmental profile is obtained by light scattering from a pure water sample at

305 nm. This has a flill-width at half-maximum (FWHM) value of about 75 ps. The

fluorescence Hfetimes ofthe exched states of l-naphthol are measured at 360 nm for neutral

l-naphthol and at 460 nm for the anionic species. Decay lifetimes are obtained by fitting

the data stored in the computer using the herative reconvolution technique. In order to judge

the quaHty ofthe fits, plots of weighted residuals and the autocorrelation functions have been

examined along with the reduced y^ value.

121

.iiUX!':',",' ' . i . . . i i i i i | i " - .


7.2. Analysis of Water Confined in Small Volumes

7.2.1. Interfacial Water in Thin Films

In this series of experiments, water containing l-naphthol was confined as a thin film

supported by a gold wire ring. Water in between thin plates and in pores can be affected by

contamination caused by the solid surface. But a system such as a thin film of water that

contains only water and the probe can be formed even without the addition of an added

surfactant. This is because the added probe, l-naphthol, hself acts as a modest surfactant

which stabilizes the film. The thinfilmof water layer is created due to surface tension.

A molecule in the center of a liquid is attracted equally in all directions by

surrounding molecules. Molecules on the surface ofthe liquid, however, are attracted only

toward the interior ofthe liquid. The surface molecules, therefore, are pulled inward, and

the surface area ofthe liquid tends to be minimized. Thefilmat the edges by theringtakes

on the same thickness as the wire. But moving towards the center ofthe film, it continually

decreases in thickness.

Thesefilmsare not very stable and do not last very long under the effect of constant

pressure and the constant exposure to laser pulses. So one must work fast to obtain a

measurement before the sample disintegrates. A relatively thin gold wire was needed to

create a film that would be stable for about 10 min. Also, relatively high concentration

solutions of approximately 1.0 x 10^ M must be used for the l-naphthol to act as a surfactant

and aid in stabilizing the film. Such concentrations are also required to obtain a measurable

fluorescence signal strength. A gold wire of diameter 0.13 mm was for this study. The

diameter ofthe film that was formed was about 0.5 cm. The gold wire ring, by itself.

122

L-:)JE:KSSSM S^
without the thin water film caused only very weak emission and this low-level interference

was subtracted from the totalfluorescencedecay measuredfromthe test sample.

7.2.1.1 Analysis ofthe Decays

The experimental data can be fit reasonably well to bi-exponential decays. Figure

7.1 shows the bi-exponential fit to the decay of l-naphthol in aflatwater film supported by

a gold wire ring. The slow decay (^690 ps) is attributed to chemically stiff or 'bound'

interfacial water. The faster decay has a lifetime of =85 ps. The faster decay is very close

to the bulk water value of 60 ps. From the data analysis h was known that the percentage

intensity of the long-lived component is much smaller (about 8%) than the short-lived

component. In an earUer study [106], the thickness of such afilmwas calculated to be about

400 A. Assuming thefilmthickness is the same in this case, the extent of highly stmctured

water near the air/water interface must be about 20 A. Chemically stiff water lies near the

air/water interface, while bulk or 'free' water exists in the middle regions ofthe film.

7.2.1.2. Computational Results Pertaining to Interfacial Water

Results of computer simulation work on interfacial water carried out in our

laboratory by Zhu et al. [178] using a modified five-she ST2 (MST-FP) [179] water model

lend support to the experimental observations made here. The interfacial MST-FP water is

123
'mm; ..'... -yjis^-.-A.._.^.^.v^.t^^ ..V.^M...^...,..,,,^^.,^..</.-.. .^..A...* .-z^....v-

0 50 100 150 200 250 300 350 400 450 500


TIME (ps) *10^

Figure 7.1. 1-Naphthol Decay lifetime in Water Confined in an Ultra-thin Film


The bi-exponential fitted data are shown by a solid line and the experimental data by dots.

124

j ^ " ' ^ ^&


an ultra-thin film forming a double-sided liquid-vapor interface [180]. The MST-FP liquid-

vapor profiles indicate a transhion from the vapor side, first to dense ice-II-type where the

bonding configurafions are more compact, then to a ice-I-type stmcture before the bulk water

properties are attained.

7.2.2. Water Confined Between Quartz Plates

Attempts were made to study water containing l-naphthol between two 1" diameter

suprasil quartz plates (supplied by RMI) whh surface flatness of A,/10. The average path

length for such a setup would be about 140 A [106]. An approximate double exponential

decay of l-naphthol in liquid water between the plates was observed. The analysis showed

that the long-lived component was about 15%. Considerable emissionfromthe quartz plates

interfered with the total fluorescence decay. Since the sample size was extremely small,

subtracting out the interference from the plates did not prove very appropriate. In spite of

the experimental uncertainties in these preliminary plate experiments, a consistent picture

seems to be emerging. In agreement with the original hypothesis, interfacial water is

orientationally stiffer than bulk water.

By using better quality LTV grade plates, such as those made whh sapphire, the

interference from the fluorescence of the plate material itself can be minimized and

characteristics of the water confined in between the plates can be determined more

accurately. Also, experiments whh different plate separations using spacers can be carried

out to explore in better detail the stmctural features ofthe various water layers. Another

series of worthwhile experiments should concern the temperature dependence ofthe decays;

125
this would give us an estimate of the orientational binding energies of water in various

interfacial regions.

7.3. Conclusions

A more stmctured and orientationally stiffer water environment has been shown to

exist near surfaces and air interfaces. Computational studies using a MST-FP water model

have also arrived at similar conclusions; the bonding configurations of water molecules near

the surface seem to be closer to the compactly arranged dense ice -Il-type. These results

imply that one cannot simply assume that water has bulk properties wherever it exists. This

would be particularly relevant in the confined regions of biological cells. Many ftindamental

biological processes occur in so smaU a volume that the water involved must be interfacial.

126

f.wi'' \M.'jii.u4Jj :
CHAPTER VIII

CONCLUSIONS

The mixture model concept presented here, which is certainly simpler than other

proposed 'liquid theory' models, is able to explain quantitatively the density maximum of

liquid H2O at 3.98 C and liquid D2O at 11.18 C and the minimum in the isothermal

compressibility at 46.5 C of Hquid water.

The reported density data of liquid HjO over the temperature range -30 to +70 C

can befitto this mixture model whh six-to seven-decimal-point precision [116]. The output

parameters from the exact algebraic fits and the least-squares fits indicate the presence of

capacious intermolecular bonding with a density extremely close to that of ordinary ice-Ih,

intermixed with compactly bonded regions having a density near that ofthe common forms

of ice, in particular ice-II. Densities at higher temperatures could also be fit to good

precision with such a model, though the model must clearly become less valid as the

temperaturerisesand more varied bonding forms begin to contribute. Thefittingprocedure

also shows that both the capacious and dense components have poshive thermal expansion

coefficients that are similar in magnitude to those of their respective ice forms. As T

approaches the vicinity of 225 K in the deep supercooled regime, the stmcture ofthe liquid

approaches disordered ice-I-type bonding, with no contribution from the deeply bonded

component.

The model is applied to the densities of liquid D2O in the temperature range -20 to

+70 C [117]. It is found that the available experimental data can be fit to the reported five-

127
decimal-point precision. The output parameters from DjO indicate the existence ofthe same

two general types of bonding characteristics in the liquid as were found in liquid HjO.

A simple 'temperature scaling law' showed that when the temperatures of isotopic

liquid water are scaled according to their temperatures of maximum density, the stmctural

properties of these liquids would very closely be the same. The idea is found to be valid for

densities of isotopic water within the precision allowed by the experimental data.

Comparing the denshies of liquid H2O and liquid DjO, the temperatures must be shifted by

7.2 C for liquid DjO and the density ratio is 1.1060 [119]. In the case of liquid T2O, the

temperature shift is 9.4 C and the density ratio is 1.2150.

In the present work, a quanthative assessment is given in terms of mixture model

parameters ofthe temperature dependence ofthe isothermal compressibility of liquid HjO

at 1 atm pressure. Published isothermal compressibility factors in the temperature range -7

to +50 C were utilized for analysis. The fitting procedure shows that the 'anomalous'

minimum in this quantity at 46.5 C emerges naturally [118].

Besides the semi-quanthative assessment ofthe dynamical data in ref 41, there are

now three extensive sets of very accurate experimental data that have been shown

quanthatively to be consistent with the mixture model analysis for liquid water: the densities

of liquid H2O and D2O and the isothermal compressibility factors for H2O. Even better

correlations should be possible when more precise and more extensive experimental

densities for DjO and T2O and compressibilities for both HjO and DjO are available.

Missing particularly are good pressure dependent experimental data at supercooled

temperatures.

128

wp^ip""*^*^^^^'"'-""""""'"^"'"" .1";'!
Aside from the good fits ofthe wide variety of experimental data on liquid water, it

is important to reiterate the fact that there already exists independent evidence that can be

used to support the proposed model. First and foremost, direct stmctural observations [123]

have been made for the types of 0 - 0 bonding expected for the suggested bulky and dense

components. The resuhs of Bosio et al. confirmed that the bonding variations between the

two configurations are not at the nearest-neighbor level but occur instead in the outlying

non-hydrogen-bonded next-nearest-neighbor 0 - 0 stmcture. Kamb analyzed the x-ray

diffraction data of Morgan and Warren [26] and reached a similar conclusion about these

mixed bonding forms in liquid water [15].

The decay lifetime of l-naphthol, a fluorescence probe, has been studied

experimentally in liquid water confined in small volumes using ultrafast laser methods. The

best results were obtained using a sample containing a thin film sample of water supported

by a gold wire ring. The experimental data ofthe decay of l-naphthol are fit by a bi-

exponential function indicating the existence of bulk water at the center of the film and

stmcturally stiffer water near the surface of the film. The slow decay is attributed to

chemically stiff interfacial water.

129
REFERENCES

I] F. Franks, in Water. A Comprehensive Treatise, Vol.2, ed. F. Franks (New York:


Plenum:, 1973).

2] J.B.F.N. Engberts, Chem.Weekbl.Mag. 491 (1979).

3] W. Kauzmann, m Advances in Protein Chemistry, Vol.XIV, eds. C.B. Anfinsen, Jr., K.


Bailey, M.L. Anson and J.T. Edsall (New York: Academic, 1959).

4] Biophysics of Water, eds. F.Franks and S.F. Mathias (New York: Wiley-Interscience,
1982).

5] J. A. McCammon and S.C. Harvey, Dynamics of Proteins and Nucleic Acids (Cambridge:
Cambridge University Press, 1987).

6] Water and Ions in Biomolecular Systems, eds. D. Vasilescu, J. Jaz, L. Packer and B.
Pullman (Basel: Birkhauser Veriag, 1990).

7] P.E. Smith and B.M. Pettitt, J.Phys.Chem. 98, 9700 (1994).

8] C.L. Brooks, IH and M. Karplus, Methods Enzymol. 127, 369 (1986).

9] Water. A Comprehensive Treatise, Vol.4, ed.F. Franks (New York: Plenum, 1975).

10] D. Eisenberg and W. Kauzmann, 77?^ Structure and Properties of Water (New York:
Oxford University Press, 1969).

II] F. van der Leeden, F.L. Troise and D.K. Todd, 77?^ Water Encyclopedia, 2nd Ed.
(Chelsea, MI: Lewis Publishers, 1990).

12] N.E. Dorsey, Properties of Ordinary Water-Substance (New York: Reinhold, 1940).

13] L. Pauling, 77?^ Chemical Bond {IXhsiCdi, NY: Cornell, 1967).

14] J.L. Kavanau, Water and Solute-Water Interactions {San Francisco: Holden-Day, 1964).

15] B. Kamb, in Structural Chemistry and Molecular Biology, eds. A. Rich and N.
Davidson (San Francisco: W.H. Freeman, 1968).

16] F.H. SmmgQv, Adv.Chem. Phys.31, 1 (1975).

130

WI!!W<BW<iWi-i'i**^i""W
17] G. S. Kell, J.Chem. Eng Data \2 , 66 (1967).

18] G. S. KeH, J.Chem. Eng. Data 20 , 97 (1975).

19] K.E. Bett and J.B. Cappi, Nature 207 , 620 (1965).

20]reflO,Chap.3.

21] W.C. Rontgen, Ann.Phy.s.u.Chem. 45 , 91 (1892).

22] P.W. Bridgman, Proc. Amer. Acad. Arts Sci. 47 , 441 (1912).

23] W. Sutheriand, Phil. Mag 505}, 460 (1900).

24] G. Tammann, Z Anorg. Allgem.Chem.lSS , 1 (1926).

25] J.D. Bemal and R.H. Fowler, J.Chem.Phys. 1,515 (1933).

26] J. Morgan and B. Warren, J.Chem.Phys. 6 , 666 (1938).

27] A.R. Ubbelohde, Melting and Crystal Stmcture (New York: Oxford University Press,
1965).

28] L.F. Evans, J. Appl. Phys., 38 , 4930 (1968).

29] R. L. McFarian, J Chem. Phys., 4, 60, 253 (1936).

30] E. R. Lippincott, C.E. Weir and A.van Valkenberg, J Chem. Phys., 32 612 (1960).

31]refl5, p.507.

32] B. Kamb, Trans. Amer. Cryst. Assoc. 5, 61 (1969).

33] B. Kamb, in Water and Aqueous Solutions, ed. R.A. Home (New York: Wiley-
Interscience, 1972) chap. 1.

34] B. Kamb, Science, 148, 232 (1965).

35] S.J. LaPlaca, W.C. Hamihon, B. Kamb and A. Prakash, J.Chem.Phys. 58, 567 (1973).

36] E. WhaHey, J.B.R. Heath and D.W. Davidson, J.Chem.Phys. 48, 2362 (1968).

37] G.A. Jeffrey and R.K. McMullen, Prog.InorgChem., 8, 43 (1967).

131

.Bi.Ill 111.niii.i.iigmf
ms^
38]ref33, p.l7.

39] V.P. Dmitriev, S.B. Rochal and P. Toledano, Phys.Rev. Lett. 71, 553 (1993).

40] D.R. MacFariane and C.A. Angell, J.Phys.Chem. 88, 759 (1984).

41] M.-P. Bassez, J. Lee and G.W. Robinson, J.Phys.Chem. 1, 5818 (1987).

42] M.-C. Belissent-Funel, J.Teixeira and L. Bosio, in Physics and Chemistry of Ice, eds.
N.Maeno and T. Hondoh (Sapporo: Hokkaido University Press, 1992).

43] L. Hall, Phys. Rev. 73, 775 (1948).

44] A. Eucken, Nachr.Ges.Wiss.Gottingen 1946. 38.

45] H.S. Frank and M.W. Evans, J.Chem.Phys. U 507 (1945).

46] K. Grjotheim and J. Krogh-Moe, Acta Chem.Scand. 8, 1193 (1954).

47] G. Wada, Bull.Chem.Soc.Japan 34, 604 (1961).

48] CM. Davis and T.A. Litovitz, J.Chem.Phys. 42, 2563 (1965).

49] A.H. Narten and H. A. Levy, Science 165, 447 (1969).

50] M.S. Jhon, J. Grosh, T. Ree and H. Eyring, J.Chem.Phys. 44, 1465 (1966).

51] H. Eyring and T. Ree, Proc. Natl. Acad Sci. (U.S.) 47, 526 (1961).

52]reflO, chap.5.

53] G. Nemethy and H.A. Scheraga, J.Chem.Phys. 36, 3382 (1962).

54] H.S. Frank and WY. Wen, Disc.Faraday Soc. 24, 133 (1957).

55] K. Buijs and G. Choppin, J.Chem.Phys. 39, 2035 (1963).

56] G. Walrafen, in Hydrogen-Bonded Solvent Systems, eds. A.K. Covington and B.P. Jones
(London: Taylor and Francis, 1968) p.9.

57] G. Walrafen, J.Chem.Phy. 47 114 (1967).

58] G. Walrafen, in Water: A Comprehensive Treatise, Vol.1, ed. F. Franks (New York
Plenum Press, 1971).

132

PIKbaia***>*Mi*IMiWBIHI^
59] M. Colles, G. Walrafen and K. Wecht, Chem.Phys.Letters 4, 621 (1970).

60] O. YA Samoilov in Structure ofAqueous Electrolyte Solutions and the Hydration of Ions,
(New York: Consultants Bureau, 1965).

61] M.D. Danford and H.A. Levy, J.Am.Chem.Soc. 84, 3965 (1962).

62] A.H. Narten, M.D. Danford and H.A. Levy, Discuss. Faraday Soc. 43, 97 (1967).

63] L. Pauling in Hydrogen Bonding, ed. D. Hadzi ( London: Pergamon Press, 1959).

64] H.S. Frank and A.S. Quist, J.Chem.Phys. 34, 604 (1961).

65] J.A. Pople, Proc.R.Soc.(London), Ser.A., 205, 163 (1951).

66] H.S. Frank, Proc.RSoc. KlAl, 481 (1958).

67]refl0, p.267.

68] J.D. Bemal, ProcKSoc. A 280, 299 (1964).

69] M. Chadwell, Chem.Revs. 4, 375 (1927).

70] F.S. Feates and D.J.G. Ives, J.Chem.Soc.(I.ondon) 1956. 2798.

71] G.W. Robinson, S.-B. Zhu, S. Singh and M.W. Evans in Water in Biology, Chemistry
and Physics (Singapore: World Scientific, 1996). in print

72] J.A. Barker, Lattice Theories ofthe Liquid State (New York: MacMillan, 1963).

73] P.D. Fleming, III and J.H. Gibbs, JStat.Phys. 10, 351 (1974).

74] G.M. BeH, JPhys. C 5, 889 (1972).

75] I. Prigogine, The Molecular Theory of Solutions (Amsterdam: North-Holland, 1957).

76] A.C. Belch and S.A. T(lce J.Chem.Phys. 86, 5676 (1987).

[77] J.K. Percus and G.J. Yevick, Phys.Rev. HO, 1 (1958).

[78] T. Morita and K. Hiroike, Prog.Theor.Phys. 23, 1003 (1960).

[79] M.S. Green, J.Chem.Phys. 33 1403 (1960).

133

,'j-'.,\ '^'..^^^^'-

-^^=j''^
[80] E. Meeron, J.Math.Phys. \_ 192 (1960).

[81] L.S. Omstein and F. Zemike, Proc. Acad. Sci. Amsterdam 17, 793 (1914).

[82] . Waisman, J.Chem.Phys. 59, 495 (1973).

[83] F. Lado, Phys.Rev. 135, Al013 (1964).

[84] T. Ichiye and A.D.J. Haymet, J.Chem.Phys. 89, 4315 (1988).

[85] L.J. Lowden and D. Chandler, J.Chem.Phys. 59, 6587 (1973).

[86] F. Hirata, B.M. Pettitt and P.J. Rossky, J.Chem.Phys. 11, 509 (1982).

[87] B.M. Pettitt and P.J. Rossky, J.Chem.Phys. 78, 7296 (1983).

[88] C.A. An%Q\\, Ann.Rev.Phys.Chem. 34, 593 (1983).

[89] R.J. Speedy, J.Phys.Chem. 96, 2322 (1992).

[90] D.W. Heerman, Computer Simulation Methods in Theoretical Physics, 2nd Ed. (Beriin:
Springer, 1990).

[91] M.P. Allen and D.J. Tildesley, Computer Simulation of Liquids (Oxford: Claredon,
1987).

[92] K. Rasmussen, Potential Energy Functions in Conformational Analysis, Lecture Notes


in Chemistry, Vol.37 (Beriin: Springer, 1985).

[93] W. Dorst-Hansen and J.S. Clegg, eds., Cell-Associated Water (New York: Academic
Press, 1979).

[94] P.J. Rossky and S.H. Lee, Chemica Scripta 29A, 93 (1989).

[95] S.-B. Zhu and G.W. Robinson, J.Chem.Phys. 94, 1403 (1991).

[96] J.J. Magda, M. TirreH and H.T. David, J.Chem.Phys. 83, 1888 (1985).

[97] M. Matsumoto and Y. Kataoka, J.Chem.Phys. 88, 3233 (1988).

[98] M.A. Wilson, A. Pohorille and L.R. Pratt, J.Chem.Phys. 88, 3281 (1988).

[99] O.A. Karim and A.D. Haymet, J.Chem.Phys.^, 6889 (1988).

134

IlliWiliil nil iiMiiWiiWi

mm
100] S.-B. Zhu, T.G. FilHngim and G.W. Robinson, J.Phys.Chem. 95, 1002 (1991).

101] S.-B. Zhu and G.W. Robinson, J.Chem.Phys. 97, 4336 (1992).

102] C.Y. Lee, J.A. McCammon and P.J. Rossky, J.Chem.Phys. 80, 4448 (1984).

103] S.-B. Zhu, J.-B.Zhu and G.W. Robinson, Phys.Rev. A 44, 2602 (1991).

104] K.J. Packer in Progress in Nuclear Magnetic Resonance Spectroscopy, eds. J.W.
Emsley, J. Feeney and L.H. Sutcliffe (Oxford: Pergamon, 1967) Vol. 3, p 87.

105] H.A. Resing, Adv.Mol. Relaxation Processes 3, 199 (1972).

106] T.G. FilHngim, S.-B. Zhu, S. Yao, J. Lee and G.W. Robinson, Chem.Phys.Lett. 161.
444(1989).

107] J.E. Hansen, . Pines and G.R. Fleming, J.Phys.Chem. 96, 6904 (1992).

108] H. Boutin, G.J. Safford and H.R. Danner, J.Chem.Phys. 39, 488 (1963).

109] P.L. Poole and J.L. Finney in Biophysics of Water, eds. F. Franks and S. Mathias
(London: Wiley Interscience, 1982).

110] Y.R. ShQn,J.Vac.Sci.TechnolBl, 1464 (1985).

I l l ] M.C. Goh, J.M. Hicks, K. Kemnitz, G.R. Pinto, K.B. Eisenthal and T.F. Heinz,
J.Phys.Chem. 92, 5074 (1988).

112] C.H. Cho, M. Chung, J. Lee, T. Nguyen, S. Singh, M. Vedamuthu, S. Yao, J.-B. Zhu
and G.W. Robinson, J.Phys.Chem. 99, 7806 (1995).

113] H. Hauser in Reverse Micelles, eds. P.L. Luisi and B.E. Straub (New York: Plenum
Press, 1984) p 37.

114] W. Kauzmann, L'Eau Syst. Biol., Colloq.Int. C.N.R.S. 246, 63 (1975).

115] For example: H. Endo, J.Chem.Phys. 72, 4324 (1980).

116] M. Vedamuthu, S. Singh and G.W. Robinson, J.Phys.Chem. 98, 2222 (1994).

117] M. Vedamuthu, S. Singh and G.W. Robinson, J.Phys.Chem. 98, 8591 (1994).

118] M. Vedamuthu, S. Singh and G.W Robinson, J.Phys.Chem. 99, 9263 (1995).

135
119] M. Vedamuthu, S. Singh and G.W. Robinson, J.Phys.Chem. 100, 3825 (1996).

120] G. Nemethy and H.A. Scheraga, J.Chem.Phys. 41, 680 (1964).

121] F. Sciortino, A. Geiger and H.E. Stanley, Phys.Rev.Lett. 65, 3452 (1990).

122] L. Bosio, S.-H. Chen and J. Teixeira, JPhys. Rev. A27, 1468 (1983).

123] S.-H. Chen and J. TeixensL, Adv.Chem.Phys. 64, 1 (1986).

124] L. Bosio, J. Teixeira and M.-C. BelHssent-Funel, Phys.Rev. A39, 6612 (1989).

125] Yu. A. Osipov, B.V. Zheleznyi, NF. Fondarenko, Russ.J.Phys.Chem. (End. Transl.)
51,748(1977).

126] R.H. Fowler, Statistical Mechanics, 2nd Ed. (Cambridge: Cambridge University Press
1966) p. 810.

127] L. Pauling, Phys.Rev. 36, 430 (1930). R.H. Fowler, ProcKSoc. London, A149. 1
(1935).

\2%]C.T>QS^xtXz, Ann.de Chim.ie et de Phys. 70, 23 (1837).

129] J.F. Mohler, Phys.Rev. 35, 236 (1912).

130] J.A. Schufle, Chem.Ind. 16, 690 (1965).

131] J.A. Schufle and M. Venugopalan, J.Geophys.Res. 72, 3271 (1967).

132] B.V. Zheleznyi, Russ.J.Phys.Chem. (Engl. Transl.) 43, 1311 (1969).

133] D.E. Hare and CM. Sorensen, J.Chem.Phys. 84, 5085 (1986).

134] For example, E.K. Bigg, Proc.Phys.Soc. B 66, 688 (1953).

135] J.V Leyendekkers and R.J. Hunter, J.Chem.Phys. 82, 1440 (1985).

136] D.E. Hare and CM. Sorensen, J.Chem.Phys. 87, 4840 (1987).

137] G.A. Baker, Jr. Essentials ofPade Approximants (New York: Academic Press, 1975).

138] see footnote to Table III in ref [18].

136

i- Wfif V I
^fBasaaBmm
m
[139] M.G. Sceats, S.A. Rice In Water. A Comprehensive Treatise, Vol.7, ed. F. Franks
(New York: Plenum, 1982). H.E. Stanley and J. Teixeira, J.Chem.Phys. 73, 3404
(1980).

140] C.A. Angdl, Annu.Rev.Phys.Chem. 34, 593 (1983).

141] Y. Xie, K.F. Ludwig, Jr., G. Morales, D.E. Hare and CM. Sorensen, Phys.Rev.Lett.
21,2050(1993).

142] E.W. Lang, D. Giriich, H.-D. Ludemann, L. Piculell and D. Muller, J.Chem.Phys. 93,
4796(1990).

143] H.E. Stanley and J. Teixeira, J.Chem.Phys. 73, 3404 (1980).

144] C.A. Angell, In Water. A Comprehensive Treatise, Vol.7, ed. F. Franks


(New York: Plenum, 1982) p.l.

145] K. Lonsdale, Proc.RSoc. London, A247, 424 (1958).

146] W.E. Thiessen and A.M. Narten, J.Chem.Phys. 77, 2656 (1982).

147] M. Goldblatt, J.Phys.Chem. 84, 5085 (1964).

148] A.R. Richards, Ind.Eng.Chem., Anal. Ed. U, 44 (1939).

149] Eq.2 of ref 127.

150] A.K. Soper and M.G. Phillips, Chem.Phys. 107, 47 (1986).

151] E.W. Lang and H.-D. Ludemann, in NMR Basic Principles and Progress, (Beriin:
Spinger Veriag, 1990).

152] Fig. 5 of ref 41.

153] H.-D. Ludemann, Ber. Bunsen-Ges.Phys.Chem. 85, 603 (1981).

154] G.S. KeH and E.WhaHey, J.Chem.Phys. 62, 3496 (1975).

155] W. A. Walker, Thermodynamic Properties of Liquid Water to One Ki lobar. NOLTR


66-217; Explosions Research Department, U.S. Naval Ordinance Laboratory, White
Oak, MD(1966).

137

Mm mm) iiiJ>i* SWC^raaaatp*


^^^^^"TTT
156] V.A. Del Grosso, JAcoust.Soc.Amer. 47, 947 (1970).

157] V.A. Del Grosso and CW. Mader, JAcoust.Soc.Amer. 52, 1442 (1972).

158] T.W. Richards and C.L. Speyers, J.Am.Chem.Soc. 36, 491 (1914).

159] G.E. Walrafen, M.R. Fisher, M.S. Hokmabadi and W.-H. Yang, J.Chem.Phys. 85,
6970(1986).

160] S.-H. Chen, K. Toukan, CK. Loong, D.L. Price and J. Teixeira, Phys.Rev.Lett. 53,
1360(1984).

161] H.R. Vruppacher, J.Chem.Phys. 56, 101 (1972).

162] E.W. Lang and H.-D. Ludemann, Ber.Bunsen-Ges.Phys.Chem. 85, 603 (1981).

163] H.-D. Ludemann and E.W. Lang, J.Phys. (Les Ulis, Fr.) Suppl. No. 9, 41 (1984).

164] J.B. Hasted, Aqueous Dielectrics (Chapman and Hall: London, 1973).

165] D. Bertolini, M. Cassettari and G. Salvetti, J.Chem.Phys. 76, 3285 (1982).

166] G. d'Arrigo, G. Maisano, F. Mallamace, P. Migliardo and F. Wanderiingh,


J.Chem.Phys. 11, 4264 (1981).

167] D.E. Hare and CM. Sorensen, J.Chem.Phys. 93, 25 (1990).

168] D.E. Hare and CM. Sorensen, J.Chem.Phys. 93, 6954 (1990).

169] E. Frank and H. Lindner, Doctoral Dissertation ofH. Lindner, University of


Karismhe, 1970.

170] R.J. Speedy and C.A. Angell, J.Chem.Phys. 65, 851 (1976).

171] C.H. Perry and D.K. Agrawal, Solid State Commun. 8, 225 (1970).

172] L. Pauling, J.Am.Chem.Soc. 57, 2680 (1935).

173] S.R. Billeter, P.M. King and W.F. van Gunsteren, J.Chem.Phys. 100, 6692 (1994).

174] A. Wallqvist and P.-O. Astrand, J.Chem.Phys. 102, 6559 (1995).

175] C.H. Cho, S. Singh and G.W. Robinson, Phys.Rev.Lett. 76, 1651 (1996).

138

f nmrr-Tlt^Sfit^^^^^^^^^^^^^^^^^^^^^^^^^^^'-'-^^^'-''-^ IJJ^^^l. . ' ^ g ^ a g '


L2M S^S
[176] H. Takahashi, Proc.Phys.-Math.Soc. Japan, 24, 60 (1942).

[177] G.W. Robinson, P.J. Thistlethwahe and J. Lee, J.Phys.Chem. 90, 4224 (1986).

[178] S.-B. Zhu, S. Singh and G.W. Ko\imson, J.Chem.Phys. 95, 2791 (1991).

[179] F.H. Stillinger and A. Rahman, J.Chem.Phys. 60, 1545 (1974).

[180] G.W. Robinson and S.-B. Zhu, in Reaction Dynamics in Clusters and Condensed
Phases, eds. J. Jortner et al., (Netheriands: Kluwer Academic Publishers, 1994) p.423.

[181] P.R. Bevington, Reduction and Error Analysis for the Physical Sciences (New York:
McGraw-Hill, 1969) chap 11.

[182] W.H. Press, B.P. Flannery, S.A. Vatteriing, Numerical Recipes in C: The Art of
Scientific Computing (Cambridge: Cambridge University Press, 1988) chap 14.

139
APPENDIX

LEAST SQUARES FITTING PROCEDURE

According to the method of least squares, the optimum values ofthe parameters aj

of an arbitrary ftinction y(x) are obtained by minimizing x with respect to each of the

parameters simultaneously. There are a number of ways offindingthe minimum value; it

can be done by searching parameter space by grid search, gradient search method or by

using approximate analytical methods.

A convenient algorithm, which behaves like a gradient search for thefirstportion of

the search and behaves more like an analytical solution as the solution converges, was

developed by Levenberg and Marquardt [181]. This gradient expansion algorithm which

combines the best features of gradient search with the method of linearizing the fitting

function can be obtained by increasing the diagonal terms ofthe curvature matrix by a factor

of A which controls the interpolation ofthe algorithm between the two extremes.

The fortran routines used are takenfromPress et al. [182]. The routines were mn in

our in-house Vax 11/730, in quadmpule precision, so as not to lose the importantfifthand

sixth decimal dighs in the final denshies because of round off at each step in the herative

procedure. The computer routine using the algorithm of Levenberg and Marquardt was

found to be remarkably stable and reliable for even six to eight parameter fits.

140

"~ -..."~-. ,%-,-i-ii7Mr>i.-iwif?Ti:-i;.---|C.)HTiiir.ir.g\^y^

You might also like