Vous êtes sur la page 1sur 27

Hydrological Sciences Journal

ISSN: 0262-6667 (Print) 2150-3435 (Online) Journal homepage: https://www.tandfonline.com/loi/thsj20

On the study of water

J. C.I. DOOGE

To cite this article: J. C.I. DOOGE (1983) On the study of water, Hydrological Sciences Journal,
28:1, 23-48, DOI: 10.1080/02626668309491141

To link to this article: https://doi.org/10.1080/02626668309491141

Published online: 24 Dec 2009.

Submit your article to this journal

Article views: 1355

Citing articles: 15 View citing articles

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=thsj20
Hydrological Sciences - Journal - des Sciences Hydrologiques, 28, 1, 3/1983

On the study of water

J.C.I. DOOGE
Engineering School, University College,
Upper Merrion Street, Dublin 2, Ireland

ABSTRACT A number of questions relative to the background


of hydrology are discussed. What does the term "water"
mean in a scientific context? Why is there water on our
planet? Why is most of it in liquid form? What does
science tell us about the molecular structure of liquid
water? Can liquid water be treated as a continuum? Is
water a Newtonian fluid? What is the scientific basis
for the equations of water movement commonly used in
hydrology? Can the various approaches to the study of
water movement be reconciled with one another? What does
the concept of water mean to the physical scientist, the
biological scientist, the social scientist and above all
to the hydrologist?

Sur l'étude de l'eau


RESUME On pose et discute ici des questions fondamentales
pour l'hydrologie. Que signifie ce terme "eau" pour la
Science? Pourquoi y a-t-il de l'eau sur notre planète?
Pourquoi la majeure partie de cette eau est-elle sous une
forme liquide? Qu'est-ce-que la Science peut nous indiquer
de la structure moléculaire de l'eau liquide? Est-ce qu'on
peut traiter de l'eau liquide comme un continuum? L'eau,
est-elle un fluide Newtonien? Quelles sont les fondations
scientifiques des équations qui décrivent le mouvement de
l'eau, et qu'on emploie si souvent en hydrologie? Est-ce
qu'on peut mettre d'accord les points de vue si variés sur
le mouvement de l'eau? Que signifie ce concept de l'eau
pour ceux qui étudient la physique, la biologie, la
sociologie et surtout l'hydrologie?

INTRODUCTION
There are many definitions of hydrology but they all centre on the
occurrence and the movement of water on our planet. Accepting that
water is an all-pervading substance on the Earth, how can we describe
its occurrence and its movement in a scientific manner? Thaïes of
Miletus (624-548 BC) is generally recognized as the first philosopher
and the first natural scientist. To Thaïes, water was the original
constituent material and the persisting substratum of all things.
This view was later modified by Empedocles and Aristotle to one that
considered water as representing only the qualities of liquidity,
mobility, wetness and coldness and thus as only one of four basic
elements. In this form, the hypothesis of water as a basic building
block in nature lasted for well over 2000 years.
While modern hydrologists are rightly concerned with the many
23
24 J.C.I. Dooge

specialist approaches to the various theoretical and practical


problems involved in the occurrence and the movement of water, there
are also a number of general background questions which should be
a matter of concern for all hydrologists. It is proper for hydrol-
ogists to ask themselves what answers are given by modern science
to the following questions:
- What does the term water mean in a scientific context?
- Why is there water on our planet?
- Why is most of it in liquid form?
- What does science tell us about the molecular structure of
liquid water?
- Can liquid water be treated as a continuum?
- Is water a Newtonian fluid?
- What is the scientific basis for the equations of water movement
commonly used in hydrology?
- Can the various approaches to the study of water movement be
reconciled with one another?
Consideration of these questions should prove not only of
interest in itself but valuable as a background to the more detailed
studies that concern hydrologists.

WHY IS THERE A HYDROSPHERE ON OUR PLANET?


It is by no means obvious why a terrestrial-type planet should possess
a hydrophere of free water (Ringwood, 1978). While the low-temper-
ature condensates falling on the Earth during its planetary formation
probably contained up to 20% of water, this water would have been
continually reduced by the metallic iron in the high temperature
condensates and the hydrogen thus formed would have escaped from the
planet's gravitational field. The small amount of volatile-rich low-
temperature condensates which survived passage through the atmosphere
and became incorporated within the solid body of the accreting planet
would also be expected to be decomposed because of the great excess
of metallic iron. Hence it might be expected that Earth would, like
its own moon and like the planet Mercury, show no traces of surviving
water. Under such circumstances, there would indeed be no science of
hydrology even for extra-terrestrial beings studying the geophysics
of the Earth from outer space.
There are many hypotheses for the sources of the present planetary
atmospheres (Pollack, 1981). These include:
(a) the accretion hypothesis that the source of water and other
atmospheric components was the inclusion of a small amount of water
of volatile compounds during planetary formation;
(b) the solar nebula hypothesis that gases might have been
captured and retained by gravity from the primordial solar nebula;
(c) the similar solar wind hypothesis of capture of some of the
solar wind that flowed past the planets over thousands of millions
of years;
(d) the comet-asteroid hypothesis of a collision between a planet
and the volatile-rich comets and asteroids.
Whichever hypothesis (or combination of hypotheses) is correct, it is
clear that the evolution of the planetary atmospheres has been a
complicated one and that the present atmospheres are not the same
On the study of water 25

as the primitive atmospheres of early planetary time.


Some properties of the present atmospheres and hydrospheres of
the four terrestrial planets (Mercury, Venus, Earth and Mars) are
shown on Table 1. It will be noted that the surface temperature of
Venus is anomalously high; this is due to its dense atmosphere. In
comparison with the Earth, Mercury is seen to have virtually no
atmosphere and the gases present at the surface are seen to have the
same composition as the solar wind. The atmosphere of Mars is seen
to be about two orders of magnitude less dense than that of Earth
and the atmosphere of Venus to be about two orders of magnitude more
dense.

Table 1 Present atmospheres and hydrospheres

/lercury Venus Earth Mars

Distance from the


sun (kmX 106} 58 108 150 228
Surface temperature 167°C 457°C 15°C -54° C
Surface pressure (bars) 1CT15 90 1 7 X 10~3
Gases in atmosphere 98% He 96% C 0 2 77% N 2 95% C 0 2
2% H 2 3% N 2 21% 02 3% N 2
Water vapour content 100 ppm 10000 ppm 300 ppm
Composition of clouds Concentrated Water Dust
sulphuric acid
Bulk of hydrosphere None Vapour Liquid Solid

Carbon dioxide and water vapour are transparent to visible


radiation but absorb infrared radiation. Hence an atmosphere that
contains these gases will trap incoming solar radiation and become
heated as a consequence. This is known as the "greenhouse effect".
This effect accounts for the fact that the general surface temper-
ature of Venus with its dense atmosphere is some 400°C higher than
it would be if the surface conditions were the same as on Mercury
and that the general surface temperature of the Earth is some 15°C
higher than it would be under the surface conditions of Mercury
(Goodye & Walker, 1972). The corresponding increase in the case of
Mars with its relatively light atmosphere is only a few degrees.
The differences between the rises in surface temperature in the
three cases can be explained by considering what happens as a
result of evaporation or sublimation of water during the evolution
of a planetary atmosphere. The phase in which water occurs is
determined by the temperature and the pressure (see Fig.l). The
primitive surface temperature of Mars (which is the farthest of the
terrestrial planets from the sun) would have been about -50°C.
Water vapour and other gases would not accumulate in such an atmos-
phere for long before the pressure would reach the saturation vapour
pressure of ice. Accordingly any additional water released into the
primitive Martian atmosphere would condense on the surface as frost
and the surface temperature would cease to rise. In the case of the
Earth, the primitive temperature (resulting from radiation balance
with the sun) before the formation of an atmosphere would be a few
26 J.C.I. Dooge

degrees below the freezing point of water. Under these conditions


the saturation vapour pressure would be almost three orders of
magnitude greater than that at which Mars would reach its frozen
equilibrium. Accordingly the hydrosphere would continue to evaporate
until a more substantial atmosphere had been built up. The final
equilibrium would be reached with a general surface temperature of
about 15°C at which the saturation vapour would be 0.01 bar giving
the water vapour content in the atmosphere as 10 000 ppm.

350
300
250
200

150
o
- 100
UJ

< 50

-50
10"4 10"3 10~2 10"' 1 10 10 2 103 10 4 105
PRESSURE ( b a r s )

Fig. 1 Phase diagram of water.

In the case of Venus, the starting temperature would be some 50°C


due to its lesser distance from the sun. Consequently a greater
pressure of water vapour would be required in order to cause con-
densation than in the case of the Earth or Mars. Calculations
indicate that the greenhouse effect caused by the greater water
vapour content of the atmosphere would be so large that condensation
could never occur, thus giving rise to a runaway greenhouse effect
(Rasool & de Bergh, 1970). Consequently, any water present in the
atmosphere of Venus would be in vapour form. In fact the amount of
water vapour in the Venusian atmosphere appears to be only about
100 ppm, indicating that the water has been removed from the atmos-
phere of Venus in some way. This can be explained by the photo-
dissociation of water vapour and the escape of hydrogen from the
planet's gravitational field over a long period of time.
Our present atmosphere is probably not the first atmosphere on
our planet. The existence of a hydrosphere in liquid form is seen
to be exceptional. It would appear that the release of water from
a bound form was only possible because of the very high temperatures
and pressures in the Earth's core and that its emergence at the
surface was only possible due to complex tectonic processes and out-
gassing. Even accepting the formation of primitive oceans and
atmospheres on the Earth, a position closer to the sun by about 5%
could have resulted in a runaway greenhouse effect and the loss of
practically all of the Earth's hydrosphere to outer space.
On the study of water 27

WHY IS OUR HYDROSPHERE LIQUID?


The phase in which water will occur at any given temperature and
pressure can be derived from its thermodynamic phase diagram which
is shown as Fig.l. Given an average atmospheric pressure of 1 bar
and an average surface temperature of 15°C, the phase diagram tells
us that water should occur in liquid form since the temperature
lies between the melting point and the boiling point for this
particular pressure. However, the question must still be asked
whether we would expect a substance like water to have a melting
point of 0°C and a boiling point of 100°C at atmospheric pressure.
In the periodic table, oxygen belongs to the first order (i.e.
the first full row) and to group VI (i.e. the sixth column) since
the oxygen atom contains eight electrons. Since atoms belonging to
the same group (i.e. the same vertical column in the periodic table)
are expected to exhibit a gradation of properties with atomic weight,
it is interesting to compare water which is oxygen hydride with the
hydrides of the other elements in group VI (sulphur, selenium,
tellurium) which combine with two atoms of hydrogen in order to fill
their outermost orbitals. The melting points and boiling points
of the resulting hydrides are shown in Table 2 in descending order
of relative molecular mass.

Table 2 Comparison of group V I anhydrides

Relative Melting Boiling


molecular point point
Compound mass (°C) (°C)

H 2 Te 129.6 -51 -4
H 2 Se 81.0 -64 -42
H2S 34.1 -82 -62
H20 18.0 0 100

It is clear from this table that, if the regular gradation of


the melting point for hydrogen telluride, hydrogen selenide and
hydrogen sulphide were to apply also to oxygen hydride, then the
melting point of oxygen hydride (i.e. water) would be about -100°C.
Similarly, if the gradation in the boiling point of the group VI
anhydrides continued through tellurium, selenium and sulphur to
oxygen, the boiling point of oxygen hydride (i.e. water) would be
about -80°C. Accordingly, were it not for the anomalously high
values of the melting point and boiling point of water, water could
only exist in vapour form at geophysical temperatures and we would
have no need for the subject of hydrology as we know it.
A complete understanding of water is not possible without a
knowledge of its anomalous behaviour. Apart from the abnormally
high melting point and the abnormally high boiling point, water
displays other anomalous properties. Water has the highest latent
heat of vaporization and the highest specific heat of any substance.
Thermal conductivity of water is the highest of any substance except
28 J.C.i. Dooge

mercury. The surface tension of water is two or three times that of


common liquids (but is one or two magnitudes lower than the surface
tension of liquid metals in which different phenomena are involved).
Water has the highest dielectric constant and hence is an almost
universal solvent.
Apart from anomalously high values of various properties, water
shows anomalous variations with temperature and pressure compared
with other liquids. Most important of these anomalies are that
water has a minimum molar volume (and hence a maximum density) at
4°C rather than at its melting point of 0°C and that the specific
heat of water has a minimum between 34°C and 35°C which is close to
the body temperature of animals.

WHY HAS WATER SUCH A HIGH MELTING POINT AND BOILING POINT?
The physical chemist explains the anomalous properties of water on
the basis of the polarity of the water molecule and of the hydrogen
bonding between water molecules. Figure 2(a) shows a simplified

Fig. 2 Structure of the water molecule.

picture of the outer 2p atomic orbitals of an oxygen atom and the Is


atomic orbitals of two hydrogen atoms (Speakman, 1966). Each of the
hydrogen atoms will have one electron in the Is orbital whose full
complement is two electrons. The oxygen will have the full comple-
ment of two electrons in the Is orbital, the full complement of two
electrons in the 2s orbital and four of the possible six electrons
in the three 2p orbitals. Assuming the p orbital normal to the page
to have its full complement of electrons, the remaining orbitals
shown in Fig.2(a) will have one electron each.
If the atoms approach each other as indicated in Fig.2(a), the
overlap of the atomic orbitals, which is possible due to the wave
nature of the phenomenon, will result in the formation of molecular
orbitals, one bonding and the other anti-bonding in each case. The
two bonding orbitals are represented in Fig.2(b). These two bonding
molecular orbitals can accommodate the four electrons with a reduct-
On the study of water 29

ion of total energy and hence with stabilization of the water mole-
cule. In fact the repulsion between the two sets of bonding elect-
rons increases the angle H-O-H from 90° to 104.52°.
It must be kept in mind that these molecular orbitals are not
deterministic paths but rather standing wave patterns of electron
density. A typical pattern of electron density is shown in Fig.2(c)
(Fletcher, 1970). For the purposes of illustrating the structure of
water it is more usual to present a more deterministic picture such
as that shown in Fig.2(d). In this representation the centre of the
oxygen nucleus is taken as the centre of the molecule and the oxygen
is considered to have a van der Waals radius (based on minimum
separation) of 1.4 A and each of the hydrogen atoms to have a radius
o
of 1.2 A. Alternatively the model could be based on the average
closest packing of water molecules and the molecule could be consid-
ered as a sphere with an oxygen nucleus at the centre and a radius
of 1.41 A.
All molecules are electrically balanced but the centres of action
of the charges may not coincide. When this coincidence does not
occur, there will be a separation of the charges and the molecule is
said to be polar. There will then be a separation of charge and the
molecule will tend to align itself with a surrounding electrostatic
field. The strength of such an effect is measured by the dipole
moment which is the product of the balanced charge and the distance
of separation.
In the case of the water molecule, oxygen is more electro-negative
than hydrogen and electrons will spend more time near the oxygen
nucleus than the hydrogen nucleus. This results in separation of the
charge on each of the 0-H bonds as shown in Fig.2(e). If the two
0-H bonds were collinear, the individual moments would cancel one
another but since they make an angle of 104.5° with each other there
is a net dipole moment in the direction of the arrow shown in Fig.2(e).
For the water molecule this dipole moment has the value of
—2 0
1.82 Debye units (m-esu x 10 ).
Most of the electron density of a water molecule is contained
within a sphere of about 1.4 A radius and accordingly we may take as
a rough model of the water molecule a spherical object of this radius
with partial charges (two positive and two negative) at four points
near its periphery. These partial charges are in a tetrahedral
relationship and make angles of about 110° with one another at the
centre as indicated in Fig.2(f).
The maximum density of water can be estimated on the basis of an
o
equivalent sphere of radius 1.4 A. The weight of an individual
molecule (m) can be obtained by dividing the gram molecular weight
(M) by Avogadro's number (N):
m =
N = 6.1 x'l023 = 2 9 " 5 x l c r 2 V a m (1)
Trigonometry tells us that for the closest packing arrangement the
number of spheres per unit volume is given by

1 _ J^ (2)
n
~ 4/2 * r 3
so that for a sphere of 1.4 A (i.e. 1.4 x 10" mm) the number of
spheres per cubic centimetre will be given by
30 J.C.I. Dooge

= 6 4 . 4 x 10" (3)
4/2 (1.4 x 1 0 " 8 ) 3
so t h a t the density i s g i v e n by
2t
P mn = (29.5 x 10 *)(64.4 x 10 2 1 ) 1.9 g cm" (4)

The actual density of ice is about 0.9 g cm - which is only just half
of the density for the closest packing. This structure of ice is
quite rigid and is shown by the fact that it requires a pressure of
over 1000 atmospheres to change the structure of type I ice as
indicated by the phase diagram of water shown as Fig.l.

WHAT IS THE STRUCTURE OF BULK LIQUID WATER?


Ice at atmospheric pressure shows a tetrahedal structure with a
spacing between the oxygen atoms of 2.76 X. This tetrahedal connect-
ion of each oxygen with four other oxygens is to be expected in view
of the tetrahedral distribution of the charges in the water molecules
shown in Fig.2(f). However, the spacing of 2.76 A is somewhat less
than the spacing of 2.80 A appropriate for two oxygen atoms in
contact. Neutron diffraction studies show that in fact there is a
hydrogen atom intervening between each two oxygens and for this
condition the spacing between oxygen atoms should be 3.2 Â which is
well in excess of the observed spacing of 2.76 A. This linking of
two atoms via a hydrogen atom is known as hydrogen bunding. Though
a weak force (the strength of the hydrogen bond is about one tenth
that of the covalent 0-H bond), hydrogen bonding is extremely import-
ant in chemistry and biochemistry. The phenomena is largely restric-
ted to compounds of oxygen, fluorine and nitrogen.

(b)

Fig. 3 Tetrahedral structure of water.

The structure of ice due to hydrogen bonding is shown in Fig.3.


Figure 3(a) has a basic tetrahedral structure with four atoms linked
with hydrogen bonds. The hydrogen atom spends 50% of the time in
each of the two positions shown in Fig.3(a) i.e. it alternates
between being linked with one oxygen atom and the other. Figures
3(b) and (c) show the linking of a number of oxygen atoms and illust-
rate the arrangement of the oxygen atoms which is described as that
On the study of water 31

of crumpled hexagonal layers.


The existence of the hydrogen bonding in addition to covalent
bonding explains a number of the anomalous properties of water such
as the high melting point , the high boiling point, the high latent
heat and the high surface tension. In all these cases extra energy
is required in order to overcome the hydrogen bonding.
The anomalies whereby the molar volume has a minimum value at
4°C and the specific heat has a minimum value at 34-35° are also
readily explained by the existence of hydrogen bonding. With the
rise in temperature above 0°C there occurs a vibrational effect (as
the result of more vigorous molecular motion) which tends to increase
the molar volume and the specific heat. Howevever, there is at the
same time a structural effect due to the breakdown of hydrogen
bonding which leads to a less open structure resulting in a smaller
molar volume and a lower specific heat at constant pressure. As the
temperature increases between 0°C and 100°C the vibrational effect
and the structural effect are superimposed and lead to the observed
variations in molar volume and specific heat.
Hydrogen bonding also accounts for the other anomalous properties
of water such as the large dielectric constant of water which prevents
ions from forming salts as readily as in another liquid and hence
makes water a good and almost universal solvent.
Quantitative prediction of water properties, including its anoma-
lous behaviour, has been attempted by models of various types.
Summaries of these attempts are available in the literature (Ives &
Lemon, 1968; H o m e , 1972; Ben Nairn, 1974; Watts & McGee, 1976). Some
of these are uniformist models in which a given structure is post-
ulated for liquid water at a given temperature but which allows for
changes in the structure at certain points in the temperature range.
Others are mixture models in which two or more types of structure
are postulated with the proportions in the mixture changing with
temperature. Mixture models are of various types: two state models,
interstitial models, and bond-breaking models.
Some of the earliest models were simple aggregate models of
polymers ( H 2 0 ) n with n = 1, 2, 4 or 8. Bernai & Fowler (1933)
proposed the first structural theory of water with three preferred
molecular arrangements: (a) an ice-like water I at temperatures
below 4°C; (b) a quartz-like structure at ordinary temperatures; (c)
an ammonia-like close packed liquid at high temperatures approaching
the critical point. A later uniformist theory due to Pople (1951)
postulates a tetrahedral-like structure but takes as the key mech-
anism the bending of hydrogen bonds without any breaking of these
bonds.
An interstitial model of liquid water was proposed by Samoilov
(1946) in which the lattice of oxygen atoms shown in Fig.3(b) and
(c) is combined with the partial occupation of the 12 vertex cavities
by non-lattice molecules. Pauling (1959) suggested that water could
be considered as a clathrate hydrate of itself allowing for
interstitial molecules.
Frank & Wen (1957) suggested a water structure based on the idea
of "flickering clusters of bonded molecules mixed with, and alter-
nating roles with, non-bonded fluid". Water has a single dielectric
relaxation time of about 1 0 - 1 1 and a common activation energy of
4.6 kcal mole - 1 for dielectric relaxation, self-diffusion, shear
32 J.C.I. Dooge

viscosity and bulk viscosity. The implication is that the same


fundamental process is involved and this is interpreted as the
formation and dissolution of molecular clusters with a half life of
the order of 10~ s. Nemethy & Scheraga (1962) posed a model with
clusters of about 25 molecules and a mixture of water molecules with
4, 3, 2, 1 or 0 hydrogen bonds, the proportion varying at different
temperatures. Another mixture model is that due to Falk & Ford
(1966). Models of the above type can represent the properties of
water including the anomalous variations with temperature and
pressure to a high degree of accuracy (Eyring & Jhon, 1969).
It can be said that to the physical chemist the water molecule is
polar i.e. non-isotropic, that water has an open structure with a
density about half of that for closest packing, that water molecules
are clustered due to hydrogen bonding and that clusters act as if
they formed and reformed on a time scale of 10" s. With models
based on these concepts the physical chemist can account for the
various anomalous properties of water. However, it must be remember-
ed that these models apply to bulk water and need modification or
replacement before being applied to water near surfaces or water in
the presence of solutes.

IS WATER A NEWTONIAN FLUID?


The structure of water considered in the last section is not taken
into account in deriving the basis of the equations of motion used
in hydraulics and hydrology. Rather is water considered as a
Newtonian fluid. This approach is based on ignoring molecular
structure and assuming the fluid to be a continuum. A constitutive
equation is derived from certain assumptions and then used together
with the principles of the balance of mass, momentum and energy to
provide a unique and stable solution to a properly-posed physical
problem. Any such constitutive equation derived on the principles
of continuum mechanics will involve material parameters which must
be determined empirically. The assumptions of continuum mechanics
are not related to those of physical chemistry. In deriving con-
stitutive equations, continuum mechanics uses a number of mathemat-
ical principles as guidelines (Truesdell, 1961; Jaunzemis, 1967;
Bear, 1972). The most important of these for our present purpose is
the principle of material indifference which states that the intrin-
sic response of a material is independent of the observer.
For any body of fluid, we can divide the stress at an interior
point into a direct stress and a shear stress by writing

T
ij = - p - 6 ij + T
ij (5)

where T-^-j represents one of the nine components of the second-order


stress tensor, p represents the hydrodynamic pressure, i^- is the
corresponding component of the second-order shear tensor. If we
define a fluid as a substance which will continue to move indefinit-
ely under a shearing force, however small, then for hydrostatic
conditions the shear tensor will be identically zero and the hydro-
dynamic pressure (p) will be identical to the thermodynamic pressure
of the equation of state so that we can write
On the study of water 33
T
ij = -PCP.eKôij (6)

where p is the mass density and 0 the thermodynamic temperature,


Equation (6) is the constitutive equation for the ideal fluid of
classical hydrodynamics and for the simple Pascalian liquid of
elementary rheology.
In order to describe the movement of real fluids it is necessary
to take account of the shearing force (T^.:) and to relate it to the
nature of the fluid and the nature of the motion involved. The
first attempt to tackle this problem is due to Newton who wrote in
his Principia (1687):
"HYPOTHESIS: that the resistance which
arises from the lack of slipperiness
on the part of the liquid, other things
being equal, is proportional to the
velocity to which the parts of the
liquid are separated from one another."
This hypothesis is generalized in modern continuum mechanics as the
assumption that the shear tensor is a linear function of the deform-
ation tensor.
The first basic assumption made is the restriction of the class
of fluids to the sub-class of pure fluids i.e. fluids without memory
in which the stresses depend only on the present rates of deformation
and present values of the thermodynamic variables and not on the
history of the motion or of the thermodynamic behaviour. Since, by
definition, a fluid continues to deform under shearing stress, the
constitutive equation cannot be related to the strain but must be
related to the rate of strain (which is equivalent to the velocity
gradient) or to higher derivatives of the strain.
The second basic assumption is now made that the stress at a given
point and a given instant depends only on conditions in the immediate
neighbourhood of the point. The stress tensor will then be a function
only of the first order velocity gradient tensor and of the present
values of the thermodynamic variables i.e.

T
ij = f ij( v rs.P.Q) (7)

where V is the velocity gradient tensor defined by

3u r /3x s (8)

which is a second-order tensor.


V r g can be divided into the symmetric deformation tensor

D =
rs i(3u r /9x s + 3u s /3x r ) (9)

and the anti-symmetric rotation tensor

W =
rs 5(3u r / 3x s ~ 3u s /3x r ) (10)

By the principle of material indifference, the constitutive equation


should be independent of pure rotation and consequently we can write
34 J.C.I. Dooge

Txi = f±ô (D rs ,p,0) (11)

Since the assumption of independence of rotation was first made by


Stokes (1845), a fluid obeying the constitutive equation derived
from equation (11) is known as a Stokesian fluid.
If the further assumption is made that the shear tensor is a
linear function of the deformation tensor i.e.

xi j C i j r s (p,0)-D r s (12)

then the resulting fluid is known as a Newtonian fluid. There is no


known fluid whose behaviour has been observed to be Stokesian and
not Newtonian (Jaunzemis, 1967; Fredrickson, 1964).
The formulation of equation (12) is still far from simple since
the coefficient C-^-jrs linking the second-order shear tensor ( T ^ J )
and the second-order deformation tensor (D r s ) is a fourth-order
tensor. In general, a fourth-order tensor contains 81 components
which may all be numerically distinct. Since the shear stress tensor
(Tjj) and the deformation tensor (D r s ) are symmetrical and therefore
contain only six independent components each, the number of compon-
ents in the fourth order linking tensor C^^g is reduced from 81 to
36. If, however, it is further assumed that the fluid is isotropic
(i.e. its properties are independent of direction), then this single
assumption is sufficient to reduce the number of independent compon-
ents of the fourth-order tensor from 36 to two. This occurs since
any fourth-order isotropic tensor can be written as

C
ijrs = X-&iyôrs + M(<5ir«5js + 6is6jr> <13>

where A and u are functions of the thermodynamic variables (p and 6)


and the remainder of the right-hand side of the equation are Kron-
ecker deltas (i.e. have the value of one for i = j and the value of
0 for i * j ) .
Substitution from equation (13) into equation (12) gives us as
the shear-deformation relationship for a Newtonian fluid

T
ij = A ' 6 ij- D rr + 2u D t j (14)

where D r r is the divergence of the velocity vector given by

D = 3ui/3x-^ + 3u2/3x2 + 3u3/3x„ (15)

and D.j is the deformation tensor defined by equation (9). Substitut-


ion from equation (14) into equation (5) gives us the relationship
between the total stress tensor ( T ^ ) and the kinematics of the
motion as

Tij = (-P + ADrr)<Sij + 2 M Dij (16)

where the material parameters A and u are functions only of the


thermodynamic variables (p and Q ) .
If the Newtonian fluid characterized by equation (16) is also
incompressible, then the divergence (D r r ) will be zero and the
constitutive equation becomes
On the study of water 35

T + D
ij = -P*ij ^ ij <17>

which contains only one material parameter, the dynamic viscosity


(u) which is a function only of the thermodynamic variables and not
of the motion of fluid.
It will be noted that the key assumption which reduced the
complexity of the general relationship between the shear tensor and
the deformation tensor was the assumption that the fluid is isotropic.
This single simplification results in a constitutive equation which
involves only two material parameters in the compressible case and
only one parameter in the incompressible case. Though there is a
shortage of critical experiments, experimental results generally
indicate that water can be assumed to be a Newtonian fluid and can
for most purposes be assumed to be incompressible.
The beneficial results of assuming that water is an isotropic
fluid are in marked contrast to the position in physical chemistry
where it is only by emphasizing the consequences of the non-isotropic
nature of the structure of water that the various properties of
water can be understood and predicted. This apparent contradiction
should not deter us either from applying non-isotropic models of
water structure to problems for which these models are appropriate
or from using a simplified equation of motion based on assumption of
isotropic conditions where this is appropriate. A complete under-
standing of water would only be possible if we were able to para-
meterize from the molecular scale, in which non-isotropic assumptions
seem necessary, to the larger scale at which this condition is
apparently unimportant; we could then evaluate the error involved
in assuming isotropic conditions and thus identify the conditions
under which non-isotropic structure of water might become important.
Meanwhile, we should accept the lesson that problems can be solved
on different scales by apparently conflicting approaches.

HOW DOES A NEWTONIAN FLUID MOVE?


For an incompressible fluid, the three equations for the conservation
of linear momentum in each of the three coordinate directions can
(with the aid of the tensor notation and the summation convention
of repeated indices) be written as

P ^i + Ui int = F. - ±- (Ti1 ) da)


J x 1J
3t 3XJ 3Xj

where F^ is the i-component of the body force and T ^ the i-component


of the surface force on the plane whose normal is in the j-direction.
For the case of an incompressible Newtonian fluid the constitutive
equation of T i • given by equation (17) can be substituted into
equation (18) to give

3UH 3u-i It - ld± + » J ^ (19)


i + Ui i. =
J P P3xj p 3x-:3xj
3t 3XJ
which is known as the Navier-Stokes equations, forming the basis for
36 J.C.I. Dooge

the study of the motion of viscous fluids in continuum mechanics.


It is interesting to ask ourselves the question: what is the
basis of the validity of the Navier-Stokes equations as a description
of the movement of water? Can they be derived from the molecular
properties of water through the kinetic theory of liquids or other-
wise? Or can they only be justified through empirical verification
of the hypotheses of continuum mechanics? The former approach is
one of great difficulty and leads to the interesting results that
the equations are valid in one dimension and in three dimensions, but
not in two dimensions (Oppenheimer & Keyes , 1972, 1973). The latter
approach, based on the scale of continuum mechanics, suffers from the
disadvantage that only about half a dozen cases are known in which
there is a closed form solution for the Navier-Stokes equations.
Even in the latter cases, it is only possible to show that the
assumption of a Newtonian fluid is sufficient for the prediction of
the observed behaviours and not that it is uniquely consistent with
these results.
It is interesting to note the manner in which the Navier-Stokes
equations were first derived. Navier (1822) considered the case of
a homogeneous incompressible fluid consisting of a large number of
molecules capable of mutual action on one another. He considered
that under static conditions the molecules would be in equilibrium
under the action of mutual forces which would depend on the spacing
of the molecules from one another. He considered that an additional
force would operate when the molecules were in motion. Navier assum-
ed that the latter force could be taken as directly proportional to
the relative velocity between the two molecules concerned and as an
unknown function of the distance between them. This force can be
written as

F = f (r) . <5u (20)

where r is the spacing between the molecules and ou their relative


velocity. By means of a complex analysis, in which he assumed that
the molecules were arranged symmetrically and that the form of the
function f(r; such that the effect would decrease rapidly with
increasing values of r, Navier derived the following form for the
viscous stress term in the equation of motion

^ - diP = Ml-^L- (21)

and calculated the value of parameter u (which is the dynamic vis-


cosity) as

u = (871/30) J f(r).r'*.dr (22)


' o
Poisson (1829) derived the compressible form of the Navier-Stokes
equations on the basis of a molecular model different to that of
Navier. Poisson postulated that the motion of a fluid could be
analysed by supposing (a) that in a small interval of time the fluid
would distort in the same way as an elastic solid giving rise to
different pressures in different directions, and (b) that in a
succeeding small interval of time the molecules would rearrange
On the study of water 37

themselves in order to make the pressure the same in all directions


as in the case of a fluid. By assuming the two processes to be
independent and reducing the period to an infinitesimal length,
Poisson obtained the compressible form of the Navier-Stokes equations.
Saint-Venant (1843) sought an expression for the interior pressure
in the fluid in motion without making any hypotheses about inter-
molecular forces. He made the hypothesis that the tangential force
on any plane passing through a point is in the direction of, and
proportional to, the principal sliding along that plane. In effect
Saint-Venant assumed that the shearing stress is given by

T = M
ij ' D iJ (23)

where D^^ is the deformation tensor defined in equation (9), the


constitutive equation for a Newtonian fluid. In this way he derived
the form of the equation of motion for an incompressible viscous
fluid originally obtained by Navier.
Stokes (1845) approached the problem in much the same form in
which it is presented in present-day texts on continuum mechanics.
Though he discusses the molecules of the fluid in his introductory
remarks, Stokes' main development is free from any assumptions on
molecular behaviour. Stokes assumed the following principle:
"That the difference between the pressure
on a plane in a given direction passing
through any point P of a fluid in motion,
and the pressure which would exist in all
directions about P if the fluid in its
neighbourhood were in a state of relative
equilibrium, depends only on the relative
motion of the fluid immediately about P;
and that the relative motion due to any
motion of rotation may be eliminated
without affecting the differences of the
pressures above mentioned."

This principle of Stokes can be expressed mathematically as

T
ij = - P-Ôij + T
ij< D rs> <24>

In his paper Stokes discusses at some length the pressures and


tangential forces called into play by the relative displacement of
the particles in the fluid, and after doing so, comments that the
same result could have been obtained by assuming the shear stress
to be a linear function of the deformation and avoiding all spec-
ulation as to the molecular constitution of the fluid.
It is significant that two of the four major early presentations
of the Navier-Stokes equations depend solely on arguments of molecular
displacement and restoring forces (Navier, 1822; Poisson, 1829), one
explicitly avoids discussion of the molecular structure and uses
only a postulated relationship between the shear tensor and the
deformation tensor (Saint-Venant, 1843), and the fourth includes
both a molecular discussion and a shear-deformation postulate
(Stokes, 1845).
38 J.C.I. Dooge

ARE THE NAVIER-STOKES EQUATIONS USEFUL IN HYDROLOGY?


Even if the Navier-Stokes equations are valid at a continuum point,
we must still ask whether they are applicable at the scales of
interest in hydraulics or in hydrology. In porous media flow, where
the velocities are low enough so that the viscous forces dominate
the inertial forces sufficiently to avoid the onset of turbulence,
the equations used in practice involve substantial parameterization
of the Navier-Stokes equations. In the case of such flow, the
accelerations on the left-hand side of the Navier-Stokes equations
as given in equation (19) are neglected compared to the other terms
of the equation so that the equation at a point is written as

M ^!iL = F - IE (25)
3x-j3xj 3x-^

In order to link the simplified form of the Navier-Stokes equations


with Darcy's law for saturated flow, it is necessary first of all to
integrate across the individual pore space to obtain
aL 2 / 3D
(26)
x x
U dx±'

where L is a length characterizing the cross-sectional scale of the


pore space and a is a dimensionless number depending on the nature
of pore space. Since Darcy's equation is written in terms not of
the mean velocity in a pore but in terms of the face velocity,
equation (26) must now be integrated for a second time across a
representative cross section of the porous medium in order to give

77 _ k i F _ IE f97,
u -"• dx^

where k^ is the intrinsic permeability of the porous medium in the


i-direction. Either probabilistic models or deterministic concept-
ual models can be used in order to link the characteristics of the
medium on a microscale with the value of the intrinsic permeability
on the macroscale. One of the best known of such relationships is
the Kozeny-Carman equation given by

k = n 3 /(l - n ) 2 . d m /180 (28)

where n is the porosity and d m is the median diameter of the solid


phase of the porous medium.
The scale on which the intrinsic permeability in equation (27)
can be considered a constant will vary with the nature of the
porous material from a relatively small scale in the case of fine
sediments to a large scale in the case of fractured rocks. When we
pass from such a representative area to the behaviour of a ground-
water reservoir on catchment scale there is a further element of
parameterization. The high degree of spatial variability and the
lack of data force us to abandon the internal description based on
the scale of equation (27) and to interpret the behaviour of the
groundwater reservoir in terms of an external description based on
On the study of water 39

global parameters of transmissibility and storativity determined by


pumping tests.
In free surface flow, the same situation in regard to multiple
levels of parameterization exists. In most cases of interest in
hydraulics and hydrology, the first parameterization is an averaging
with respect to time which is needed because of the turbulent nature
of the flow. By dividing the velocity at a point into its mean value
(U) and a turbulent fluctuation (u'):

(29)

and then performing Reynolds averaging over the turbulent time scale,
we obtain the Reynolds equation of motion as

8ui dux 3(ui'ui') 3P 82ui ,„_ N


p i. + p U ^ 1 + p ± i_ = F± - —^ - u ±— (30)
3t 9xj 3XJ dx^ 3xj3xj
which includes an additional term containing the cross-product of the
turbulent fluctuations of the velocity components. In virtually all
practical problems of free surface flow, the viscous stresses
(represented by the third term on the right-hand side of equation
(30)) which reflect the constitutive equation of water as a Newtonian
fluid are negligible compared with the Reynolds "stresses" represent-
ed by the third term on the left-hand side of the equation.
An additional equation linking the new term for the turbulent
Reynolds "stresses" to the mean variable of flow is required in order
to obtain a compatible set of equations capable of solution. The
most important types of models proposed for this closure equation
have been based on eddy viscosity (Boussinesq, 1877), mixing length
(Prandtl, 1925; von Karman, 1931), transport of kinetic energy of
turbulence (Kolmogorov, 1942; Prandtl, 1945) or more complex models
involving more than one additional equation and capable only of
solution on large computers (Spaulding & Launder, 1972; Rodi, 1980).
In dealing with such hydrological problems as unsteady flow in
an open channel, either further parameterization is necessary or else
the problem is formulated directly on a macroscale. Saint-Venant
(1871) considered the motion of a slice of water filling the whole
cross section of a prismatic channel but of infinitesimal thickness
and wrote down an expression for the acceleration of this slice
under the action of gravity and of boundary shear, thus obtaining
the expression

3Û - 3u To 3z „ ,„,,
^ r + u ^ - + -r-+g^- = 0 (31)
3t 3x pR dx
in which x Q is the average boundary shear, R the hydraulic radius and
z the elevation of the water surface. Because this is a global or
external description the internal forces have cancelled out, leaving
only the shearing forces around the boundary to be accounted for.
It is still necessary for closure purposes to express the average
boundary shear in terms of the mean variables of flow. For the case
of rough turbulent flow, the boundary shear is taken as proportional
to the square of the velocity at some small representative distance
40 J.C.I. Dooge

from the boundary. If the velocity distribution of the cross section


is taken as uniform then the Chézy friction law results and if the
velocity distribution is taken as a one-sixth power law then the
Gauckler-Manning-Strickler formula results.
It is clear that the physical equations used in what are sometimes
called causal models of hydrological processes involve several levels
of parameterization of the Navier-Stokes equations derived by cont-
inuum mechanics, which in turn ignore the molecular structure of
water as established in physical chemistry. This poses a dilemma
for the hydrologist: whether to attempt a series of parameterizations
from the scale of continuum mechanics in order to establish the laws
of water movement at hydrological scale, or whether to postulate
hypotheses describing water movement at the hydrological scale and
seek to verify these hypotheses by carefully designed critical exp-
eriments .

WHAT DOES WATER MEAN TO A PHYSICAL SCIENTIST?


There is no single meaning of water and no universal model for water
even in the physical sciences. To the physical chemist, water is a
polar molecule and therefore non-isotropic. Its anomalous proper-
ties are considered capable of explanation on the basis of hydrogen
bonding. To the scientist interested in continuum mechanics, water
is considered as an isotropic fluid without memory whose constitutive
equation can be assumed to be linear and consequently whose motion
can be described by the Navier-Stokes equations. To the expert in
hydraulics, water is an incompressible fluid of turbulent behaviour
for which the Navier-Stokes equations may still be valid at a point
but which are no longer useful because of the complexity of the flow
pattern. To the geomorphologist, water is an agent of denudation
whose action has moulded the landscape and determined the drainage
pattern whether permanent or ephemeral.
At a global scale, water to the geochemist is a reactive trans-
porter in the various geochemical cycles of carbon, nitrogen,
phosphorus and sulphur. To the geophysicist, it is a climate
modifier whose anomalously high latent heat of vaporization enables
large quantities of energy to be transported from the tropics to the
polar regions through the general circulation of the atmosphere.
It is becoming increasingly clear that the differences in the
concepts of water and of the models used in the physical sciences
to simulate water depend greatly on the question of scale. An
attempt is made in Table 3 to give an approximate estimate of the
significant length and time scales for the various approaches to
the study of water. The physical chemist is interested in the
behaviour of the water molecule in which the spacing between the
o —10
oxygen and the hydrogen atoms is 0.96 A, i.e. about 10 m. Many
of the models most successful in explaining the properties of water
are based on clusters of water molecules on a rather open hexagonal
spacing so that the significant length is about 10~ m. In many of
the recent models it is assumed that such water clusters are
continually forming, breaking down and reforming with a half life
of about 1 0 " u s.
Fluid mechanics ignores the fine structure of matter and treats
On the study of water 41

Table 3 Significant length and time scales

Length (m) Time (s)

Water molecule i<r 10 10"


Water cluster 10~8 10"
Continuum point 1Û™ 5

Turbulent flow 10"2


Experimental plot 10 10
Basin module 102 10 2 -10 3
Sub-basin 103 103
Basin 10 4 -10 5 104
General circulation model 10s 10 6

water and other fluids as continua on the assumption that at the


scale of interest one can neglect the effect of the molecular behav-
iour. It is usually assumed in continuum mechanics that such an
assumption will be valid for phenomena whose length scale does not
fall below about 10" m (Batchelor, 1967). In most problems of
practical interest, the flow is turbulent and the Navier-Stokes
equations of continuum mechanics are no longer applicable to the
mean motion. Recourse is had to the Reynolds equations obtained
by averaging the turbulent fluctuations and by invoking some tur-
bulence assumption in order to ensure closure of the problem.
Whether the latter is postulated in terms of a mixing length (Prandtl,
1925) or of some more modern formulation (Rodi , 1980), there is an
inherent length scale which varies with the problem but which for
_2
many problems may be taken as about 10 m.
A variety of scales is also encountered when we attempt to des-
cribe hydrological phenomena. If we take the meaning of a signif-
icant length scale as a length below which phenomena can be neglected,
then in the case of field measurements and experimental plots the
significant length scale is about 1 m. For a finite difference grid
used to compute a physically-based model of some basin module, the
significant length scale might be taken as 10 m. Further significant
lengths can be attached to a sub-basin, to a small basin and to a
large basin. Above these hydrological scales are larger scales used
in modelling the general circulation of the atmosphere (GARP, 1975).
Here the scales of interest are 100 km (i.e. 10 m) and 10-15 days
(i.e. about 10 5 s ) .
It is clear from this discussion that the scales of interest to
the hydrologist are macroscales which are about 10 D larger than the
12
mesoscales of continuum mechanics and about 10 larger than the
microscales of physical chemistry.

WHAT DOES WATER MEAN TO A BIOLOGICAL SCIENTIST?


A clear understanding of the behaviour of water is essential for
the understanding of life processes. Equally, a clear understanding
of biological principles is essential in hydrology and water resources
42 J.C.I. Dooge

development. The following is an apt description of the link between


the physical properties of water and the nature of biological
processes (Kirk & Othmer, 1970):
"Water is a necessary condition of life. Life
began in the oceans of Earth and evolved
there. As a consequence the peculiarities of
water have been utilized by and incorporated
into the fabric and workings of the life
processes in the most delicate and intricate
way. Water to a considerable extent deter-
mines not only the structural configuration,
but also the biological function of bio-
macromolecules."
Biopolymers can only be maintained in the presence of a critical
amount of water which varies with the polymer species and the very
term protein, refers to a system of polypeptide plus water.
All the indications are that water in a cellular system is not
just an inert solvent but plays a definite role in which its special
structure is of importance (Tait & Franks, 1971; Franks, 1972;
Robinson, 1978). Just over half of the water in the human body is
intracellular water rich in potassium divided among 10 1 cells. Each
of these cells has a diameter of about 10" m and for such a size
equilibrium can be reached by diffusion in about 0.05 s. Such
cellular water is unlike bulk water in its properties but not a
great deal is known about these special properties.
To the botanist, water is a transporter of plant nutrient and a
filler material which maintains the rigidity of the cellulose walls
of the plant cells. A number of the anomalous properties of water
are of importance to the botanist. The anomalously high surface
tension gives as high as possible radius of curvature in the soil
water and therefore the greatest possible soil moisture storage for
a given matric potential. The anomalously high latent heat of
fusion is a protection against freezing and the high dielectric
constant brings ionic compounds readily into solution.
To the zoologist the extra-cellular water in the animal body
provides an internal environment which is thermostatically controlled
and chemically stable and within which the animal cells can live and
thrive. For an average human being of 70 kg weight, the intracell-
ular water rich in potassium constitutes about 23 1 and extra cell-
ular water rich in sodium constitutes about 19 1 (Robinson, 1978).
To the epidemiologist water is important as a habitat for a disease
vector and he will be concerned with (a) water-borne faecal-oral
diseases, (b) with helminths as intermediate aquatic hosts which
form part of the complex of life cycle of a parasite cycle and (c)
with water-related insect vectors which transmit disease.

WHAT DOES WATER MEAN TO THE SOCIAL SCIENTIST?


The role of the social sciences in regard to water problems has been
increasingly recognized in recent years (Kneese & Smith, 1966;
James, 1974; Field et al., 1974). The social scientist and the
sociologist are interested in the meaning attached to water by social
On the study of water 43

groups and by individuals. To them water is an important social


factor and may be defined as a means of subsistence or as a means of
communication or as a factor in technology or as a symbol in ritual
(Burch & Check, 1974).
Anthropologists and sociologists are very interested in the
effect of water at all times of human history but most notably in
regard to prehistoric cultures, the hydraulic societies of the
ancient empires of the Middle East, the significance of water in
modern times in tribal and peasant societies and the management of
water resources in developed societies (Bennet, 1974). Sociologists
are concerned with the behaviour of people in groups and the relation-
ships of personal and institution developed among men as they deal
with water problems (Johnson, 1974).
At one time considered as a non-marketable human necessity, water
is now regarded as a resource to be allocated and a large literature
has developed from the economics of such allocations (Hart, 1974,
is but one example). Thus, to the economist water is the resource
to be allocated among different users, among different periods of
time and among different regions. Indeed it may be said that not
only has economic analysis contributed to water resources develop-
ment but the problems of water resources development have given rise
to developments in economic analysis such as cost-benefit analysis
and planned programming budgeting (Hart, 1974).
To the political scientist concerned with the decision methods of
individual groups, a literature dealing with water resources develop-
ment has developed more slowly than in economics. To the social
scientist the important distinctions may be betweer marketable
irrigation water and non-marketable flood water; or may be concerned
with whether water is a suitable subject for allocation by market
prices either by negotiation among leaders of political groups, or
by hierarchical control of political leaders or by democratic control
through political representation or a referendum (Hart, 1974).

WHAT SHOULD WATER MEAN TO A HYDROLOGIST?


What then is water to the hydrologist? Is it merely a substance
which precipitates, infiltrates, percolates, flows, transpires,
evaporates, freezes, thaws, ablates, erodes, transports, deposits,
dissolves, reacts, and so on? Or does scientific hydrology need to
take the various concepts of water from different disciplines and
apply them separately to problems of different types? Or does
scientific hydrology need to fuse these differing concepts of water
so as to provide a scientific basis for hydrology? Or does scient-
ific hydrology need to develop a completely new concept of water on
a macroscale?
We have discovered that water can mean many things to scientists
in different disciplines. Firstly, the scientific concept of water
will vary between the physical sciences and the biological sciences
and the social sciences. None of these can be ignored because they
all impinge on the work of the hydrologists from time to time.
Secondly, even within one group of sciences (physical, biological
or social) there will be many concepts depending on the discipline
of the scientist and of the nature of the problem being discussed.
44 J.C.I. Dooge
There are thus many problems of communication and inter-communication
to be solved if the hydrologist is to be able to draw on all the
sciences relevant to the occurrence and movement of water on our
planet.
The past few decades have seen distinct rivalry between the
approach to hydrological problems based on physical equations and
the approach based on overall systems viewpoint. The physical
approach starts with an internal description of the hydrological
process based on conditions at a continuum point and seeks to inte-
grate this accurate picture of local conditions throughout the whole
region of interest. In contrast the systems approach takes an ext-
ernal view of the hydrological system and seeks by observation of
the input and output to formulate a mathematical description of the
system behaviour. Just as the molecular model of water behaviour
and the continuum model may contradict one another and still be
useful each for its particular problem, so too the physical approach
and the systems approach to hydrological processes can both be
extremely useful at an appropriate scale. The problem arises at the
intermediate scale in which each of the approaches may be equally
valid.
It may be that the fundamental laws of hydrology are still to be
discovered and that this will be achieved by a combination of the
parameterization of internal descriptions based on hydraulics to the
larger scales appropriate in hydrology and a careful calibration of
external descriptions of hydrological systems. An interaction
between the two approaches may result in a corpus of knowledge which
is intrinsically hydrological and not borrowed from some other branch
of science.
Examples are available of special problems in hydrology for which
the two approaches have been harmonized. This may be readily illus-
trated for the case of groundwater outflow. The classical equation
for horizontal unsteady flow in porous media is the Boussinesq
(1877) equation:

f- (Khfi)+ r(x,t) = f || (32)


dx \ dx / dt
where h is the depth of flow, K is the saturated hydraulic conduct-
ivity, r is the recharge at the surface of the water table and f is
the storativity. One external description commonly used in classic-
al hydrology is to represent groundwater outflow by an exponential
recession curve i.e. to assume that the groundwater storage acts as
a single linear reservoir. What is the relationship between these
two approaches?
The above form of the Boussinesq equation involves the simplifi-
cation from three-dimensional to one-dimensional flow. As a result
of this simplification the predicted water profile will be in error.
However, it is most interesting to note that Charnyi (1951) has shown
that for steady groundwater flow between two free surfaces the one-
dimensional solution (while it gives an incorrect groundwater
profile), gives the same flow as the full two-dimensional solution.
Hence the values of flow and change of storage which are of interest
to the hydrologist will be correctly predicted by the simplified
model.
On the study of water 45

The nonlinear Boussinesq equation can be linearized in a number of


ways. If it is linearized in ter.ns of h, the profile for the steady
state will be a parabola; if it is linearized in terms of h the profile
for the steady state will be an ellipse. The case of a groundwater
reservoir drained by parallel ditches or field drains can be used to
derive a model of groundwater outflow. Equation (32) can be solved
for this particular case subject to a delta function input, i.e. the
initial condition of a level water table with the elevation of the
water table above the impervious layer equal to ( h Q ) . The solution
which was obtained by Glover & Bittinger (1959) using linearization
in h and the separation of variable technique is given by

h - d 4 VOQ x . ntrx n27T2Kh


- > =1 . - s m — - exp
r — — 2 1 (33)
hQ - d TT ^n=l,3,...n S fS'

In the above solution h is the elevation of the water table above


the impervious layer, d is the elevation of the water surface in the
trench (or above the field drain) above the impervious layer, h Q is
the maximum (i.e. the initial elevation of the water table), x is the
horizontal distance from the trench or drain, S is the spacing of
the trenches or drains, K is the saturated permeability of the soil,
f is the drainable pore space and t is the time elapsed since the
start of the recession. If we adopt the linearization in h instead
of linearization in h, a similar equation is obtained except that it
2
will be in terms of h rather than h.
The difficulty about conflicting predictions of the shape of the
groundwater profile due to different methods of linearization does
not affect us in our study of the recession of groundwater outflow.
The outflow to a trench or a drain is given by evaluating the dis-
charge at x = 0 and x = S and combining the two values to give the
total outflow.
Using either method of linearization we obtain for the outflow
the same expression:
q = {[8kh(h - d)]/S} I™ exp[ (-n27T2KÎT/f S 2 ) . t ] (34)
n-i,j,...

Kraijenhoff van de Leur (1958) pointed out that the soil and drainage
characteristics in equation (33) are grouped into a single parameter
which he defined as the reservoir coefficient (j) so that Glover's
solution can be written as

q = {[8Kh(hQ - d)]/S} ^ n = 1 3 exp(-n2.t/j) (35)

Kraijenhoff showed that for an instantaneous input of unit volume,


we have as the impulse response

h(t) = (8/ir 2 )(l/j)f = 3)i]_ exp(-n2.t/j) (36)

Obviously, when t becomes moderately large, the first term in this


infinite series will dominate and the shape of the recession curve
will approximate that from a single linear reservoir. For small
46 J.C.I. Dooge
values of t, however, the contribution of the other terms cannot be
neglected and as t approaches zero they each approach unity and
their sum approaches infinite value. Equation (36) above represents
a one-parameter conceptual model and this is capable of being
easily fitted to field data. If the solution of equation (36) is
approximated by the first term of the series, then we have a model
corresponding to the assumption of exponential recession so widely
used in applied hydrology.
Similar comparisons can be made between the internal and external
approaches to such problems as unsteady flow in an open channel
(Dooge, 1980) but space precludes their being dealt with in this
paper. Further studies along these lines should prove fruitful in
building a coherent corpus of hydrological theory.
It is important that, as well as the various concepts of the
physical sciences, the concepts of the social scientists be fully
taken account of in the application of hydrology to the human problem.
The problem of water resources has been described by Gilbert White
(1966) as
"The fundamental problem of how man in the
face of diverse cultural tradition, social
rigidity and resource disparity, manages
peacefully to gain a more fruitful living
from the Earth."
If, as a discipline or as a vocation, hydrology is to be useful to
mankind, then this objective must affect crucially the way in which
hydrology as a science and water resources development as a tech-
nology will go about their respective tasks.

REFERENCES
Batchelor, G.K. (1967) An Introduction to Fluid Dynamics. Cambridge
University Press.
Bear, J. (1972) Dynamics of Fluids in Porous Media. Elsevier.
Ben Nairn, A. (1974) Water and Aqueous Solutions. Plenum Press.
Bernai, J.D. & Fowler, R.H. (1933) A theory of water and ionic
solutions with particular reference to hydrogen and hydroxyl ions.
J. Chem. Phys. 1, p.515.
Bennet, J.W. (1974) Anthropological contributions to the cultural
ecology and management of water resources. Chapter 2 in:
Man and Water (ed. by L.D.James), 34-81. University of Kentucky
Press.
Boussinesq, J. (1877) Essai sur la théorie des aux courants. Mêm.
Près, par div. savants à l'Acad. Sci . , Paris 23 (1) 1-680.
Burch, W.R. & Check, N.H, (1974) Social meaning of water, patterns
of variation. Chapter 3 in: Water and Community Development
(ed. by D.R.Field, J.C.Barron & B.F.Long). Ann Arbor Science.
Charnyi, I.A. (1951) Rigorous proof of the Dupuit formula for
unconfined flow. Dokl. Akad. SSR 79.
Dooge, J.C.I. (1980) Postgraduate lectures on Flow in open channels.
Department of Civil Engineering, University College, Dublin.
Eyring, H. & Jhon, M.S. (1969) Significant Liquid Structures.
Wiley.
On the study of water 47

Falk, M. & Ford, T.A. (1966) Infra-red spectrum and structure of


liquid water. Can. J. Chem. 44, p.1699.
Field, D.R. , Barron, J.C. & Long, B.F. (editors) (1974) Water and
Community Development. Ann Arbor Science.
Fletcher, N.H. (1970) The Chemical Physics of Ice. Cambridge Univer-
sity Press.
Frank, H.S. & Wen, W-Y. (1957) Structural aspects of ion-solvent
interactions in aqueous solutions; asuggestive picture of water
structure. Discuss. Faraday Soc. 24, p.133.
Franks, F. (1972) Water, the unique chemical. Chapter 1, vol. 1 of:
Water: A Comprehensive Treatise (ed. by F.Franks). Plenum Press.
Fredrickson, A.G. (1964) Principles and Applications of Rheology.
Prentice Hall.
GARP (1975) The Physical Basis of Climate and Climate Modelling.
Global Atmospheric Programme Publication series no.16 , ICSU/WMO
Geneva.
Glover, R.E. & Bittinger, M.W. (1959) Source of materials for a
course in transient groundwater hydraulics. Colorado State Univer-
sity.
Goodye, R.H. & Walker, J.C. (1972) Atmospheres. Prentice-Hall.
Hart, H.C. (1974) Towards a political science of water resources
decisions. Chapter 5 in: Man and Water (ed. by L.D.James).
University of Kentucky Press.
H o m e , R.A. (1972) Water and Aqueous Solutions. Wiley.
Ives, D.J.G. & Lemon, T.H. (1968) Structure and properties of water.
Roy. Inst. Chem. Rev. 1, 62-105.
James, L.D. (editor) (1974) Man and Water. University of Kentucky
Press.
Jaunzemis, W. (1967) Continuum Mechanics. The Macmillan Company,
New York.
Johnson, S. (1974) Recent sociological contributions to water
resource management development. Chapter 6 in: Man and Water
(ed. by L.D.James), 164-199. University of Kentucky Press.
von Karman, Th. (1930) Mechanische Ahnlicheite und Turbulenz.
Nachrichten der Akademie der Wissenschaften Gottingen Math. Phys.
Klasse 58.
Kirk, R.E. & Othmer, D.F. (1970) Article on "water properties" in:
Encyclopedia of Chemical Technology, vol. 21 (2nd edn). Inter-
science, New York.
Kneese, A.V. & Smith S.C. (1966) Water Research. John Hopkins Press.
Kolmogorov, A.N. (1942) Equations of turbulent motion of an incom-
pressible fluid. Izvestia Akad. Nauk SSSR. ser. f iz ., 6 (1-2),
56-58.
Kraijenhoff van de Leur, D.A. (1958) A study of non-steady ground-
water flow with special reference to a groundwater coefficient.
Ingénieur 17 (19), p. 87.
Navier, C. (1822) Mémoire sur les lois du mouvement des fluides.
Mêm. Acad. R. Sér de l'Inst. France 6, 389-440.
Nemethy, G. & Scheraga, H.A. (1962) Structure of water and
hydrophobic bonding in proteins. J. Chem Phys. 36, 3382.
Newton, I. (1687) Principia Mathematica Philosophiae Natural is.
Oppenheimer, I. & Keyes, T. (1972) Bilinear contributions to
equilibrium correlation functions. Phys Rev. A7, p.1384.
Oppenheimer, I. & Keyes, T. (1973) Bilinear hydrodynamics and the
48 J.C.I. Dooge

Stokes-Einstein law. Phys. Rev. A8 , p.100.


Pauling, L. (1959) Chapter 12 in: Hydrogen Bonding (ed. by L.Hadzi).
Pergamon, London.
Poisson, S.D. (1829) Mémoire sur les lois du mouvement des fluides.
Journal de l'Ecole Polytechnique XIII, can. 20, p.139.
Pollack, J.B. (1981) Atmospheres of the terrestrial planet. Chapter
6 in: The New Solar System (ed. by J.K.Beatty, B.O'Leary &
A.Chiakin), 57-70. Cambridge University Press.
Pople, J.A. (1951) Molecular association in liquids. II. A theory
of the structure of water. Proc. Roy. Soc. A205, p.163.
Prandtl, L. (1925) Bricht uber untersuchungen nur ausgebiedter
turbulenz. Z. Angew. Mat. Mech. 5, p.136.
Prandtl, L. (1945) Uber ein neues Formelsystem fur die ausgebiedter
turbulenz. Nachr. Ges. Wiss. Gottingen. Math.-Phys. Kl, 6-19.
Rasool, S.I. & de Bergh, C. (1970) The runaway greenhouse and the
accumulation of C 0 2 in the Venus atmosphere. Nature 226, 1037-
1039.
Ringwood, A.E. (1978) Water in the solar system. In: Water: Planets,
Plants and People (ed. by A.K.McIntyre) , 18-34. Australian
Academy of Sciences, Canberra.
Robinson, J.R. (1978) Water in the animal body. In: Water: Planets,
Plants and People (ed. by A.K.Mclntyre) , 60-70. Australian
Academy of Sciences, Canberra.
Rodi, W. (1980) Turbulence Models and Their Application in Hydraulics.
International Association for Hydraulic Research.
Saint-Venant, B. de (1843) Note à joindre au Mémoire sur la dynamique
des fluides. C. R. Acad. Sci., Paris 17, 1240-1243.
Saint-Venant, B. de (1871) Théorie du mouvement non-permanent des
eaux. C. R. Acad. Sci., Paris 73, 148-154 and 237-240.
Samoilov, O.Ya. (1946) Koordinatsionnal chislo v strukturu nekotorikh
zhidkostei. Zh. Fiz. Khim. 20, p.1411.
Spaulding, D.B. & Launder, B.E. (1972) Mathematical Models of
Turbulence. Academic Press.
Speakman, J.C. (1966) Molecules. McGraw-Hill, New York.
Stokes, G.G. (1845) On the theories of the internal friction of
fluids in motion and of the equilibrium of motion of elastic
bodies. Trans. Cambridge Phil. Soc. VIII, p.287.
Tait, M.J. & Franks, F. (1971) Water in biological systems. Nature
230, March 12.
Truesdell, C. (1961) Principles of Continuum Mechanics. Colloquium
Lectures in Pure and Applied Science no. 5, Sacony Mobil Company.
Watts, R.O. & McGee, I.J. (1976) Liquid State Chemical Physics.
Wiley.
White, G. (1966) Optimal flood damage management: retrospect and
prospect. In: Water Research (ed. by A.V.Kneese & S.E.Smith),
251-269. John Hopkins Press.

Vous aimerez peut-être aussi