You are on page 1of 18

Hydrogeology in North America: past and future

T. N. Narasimhan

Abstract This paper is a retrospective on the evolution of


hydrogeology in North America over the past two centuries, and a brief speculation of its future. The history of
hydrogeology is marked by developments in many different fields such as groundwater hydrology, soil mechanics, soil science, economic geology, petroleum engineering, structural geology, geochemistry, geophysics,
marine geology, and more recently, ecology. The field has
been enriched by the contributions of distinguished researchers from all these fields. At present, hydrogeology
is in transition from a state of discovering new resources
and exploiting them efficiently for maximum benefit, to
one of judicious management of finite, interconnected
resources that are vital for the sustenance of humans and
other living things. The future of hydrogeology is likely to
be dictated by the subtle balance with which the hydrological, erosional, and nutritional cycles function, and the
decision of a technological society to either adapt to the
constraints imposed by the balance, or to continue to
exploit hydrogeological systems for maximum benefit.
Although there is now a trend towards ecological and
environmental awareness, human attitudes could change
should large parts of the populated world be subjected to
the stresses of droughts that last for many decades.
Rsum Cet article est une rtrospective de lvolution
de lhydrogologie en Amrique du Nord sur les deux
derniers sicles, et une brve valuation de son futur.
Lhistoire de lhydrogologie est marque par le dveloppement de plusieurs techniques de terrain telles, lhydrologie des eaux souterraines, la mcanique des sols, les
sciences du sol, la gologie conomique, l ingnierie
ptrolire, la gologie structurale, la gochimie, la goReceived: 1 March 2004 / Accepted: 25 November 2004
Published online: 25 February 2005
 Springer-Verlag 2005
T. N. Narasimhan ())
Department of Materials Science and Engineering,
Department of Environmental Science, Policy and Management,
University of California,
210 Hearst Memorial Mining Building, Berkeley, CA,
94720-1760, USA
e-mail: tnnarasimhan@lbl.gov
Hydrogeol J (2005) 13:724

physique, la gologie marine et plus rcemment lcologie. La science a t enrichie par la contribution de plusieurs chercheurs distingus, provenant de toutes ces
branches. A prsent, lhydrogologie est  la transition
entre la volont de dcouvrir de nouvelles ressources et l
exploitation la plus bnfique au possible, et un management judicieux des ressources finies, interconnectes,
qui sont vitales pour l approvisionnement des hommes et
autres formes de vie. Le futur de l hydrogologie sera
dict par la balance subtile dans laquelle intervient les
cycles de lhydrologie, de lrosion, de la nutrition, et la
dcision dune socit technologique qui sadapterait aux
contraintes de la balance, ou qui continuerait dexploiter
les systmes hydrologiques pour un bnfice maximum.
Par ailleurs il y a une nette tendance  inclure les aspects
cologiques, les aspects environnementaux, et les changements humains qui pourraient tre influencs par les
modifications hydrogologiques observes depuis une
dizaine dannes.
Resumen Este articulo es una retrospectiva sobre la
evolucin de la hidrogeologa en Norte Amrica en los
pasados dos siglos, y una breve especulacin de su futuro.
La historia de la hidrogeologa est marcada por desarrollos en muchos campos diferentes tal como hidrologa de
aguas subterrneas, mecnica de suelos, ciencia del suelo,
geologa econmica, ingeniera del petrleo, geologa estructural, geoqumica, geofsica, geologa marina, y ms
recientemente, ecologa. El campo se ha enriquecido por
las contribuciones de investigadores distinguidos en todos
esos campos. Actualmente, la hidrogeologa se encuentra
en transicin de un estado de descubrir nuevos recursos y
explotarlos eficientemente para un beneficio mximo, a un
estado de gestin juiciosa de recursos finitos, interconectados, que son vitales para el sustento de humanos y otras
cosas vivientes. El futuro de la hidrogeologa posiblemente
est determinado por el balance sutil con el cual funcionan
los ciclos nutricionales, erosionales e hidrolgicos, y la
decisin de una sociedad tecnolgica para ya sea adaptarse
a las restricciones impuestas por el balance o para continuar con la explotacin de los sistemas hidrogeolgicos
para un beneficio mximo. Aunque existe actualmente una
tendencia hacia la conciencia ambiental y ecolgica, las
actitudes humanas podran cambiar en caso de que grandes
partes del mundo poblado estn sujetas a las presiones de
sequas que duran por muchas dcadas.
DOI 10.1007/s10040-004-0422-5

Keywords General hydrogeology History of


hydrogeology Future of hydrogeology

The waters which are from heaven, and which flow after
being dug, and even those that spring by themselves, the
bright pure waters which lead to the sea, may those divine
waters protect me here
Hymn from the Rg Veda, VII.49.2

Introduction
Lucas (1877) defined hydrogeology as Hydrogeology ...
takes up the history of rainwater from the time it leaves
the domain of the meteorologist, and investigates the
conditions under which it exists in passing through the
various rock types it percolates after leaving the surface
(Davis 1989; Mather 2001). Essentially, hydrogeology
deals with geological processes influenced by water. The
early history of hydrogeology before the 19th century was
much influenced by events in Europe. From the late 19th
century, the United States witnessed hydrogeological
developments that rivaled those of Europe. The present
work is restricted to these developments. No effort has
been made to survey the substantial hydrogeological
contributions from Europe, Russia, and elsewhere over
the past century and a half. Table 1 summarizes some of
the important events in the history of hydrogeology.
This work is unabashedly a personal statement, based
on the authors experiences and his interactions with peers
from groundwater hydrology, soil mechanics, petroleum
engineering, geomorphology, economic geology, soil
science, geochemistry, agronomy, civil engineering,
structural geology, geophysics, marine geology, history,
and law. It is unwise to believe that a single person can
have an in-depth understanding of all these fields. Imperfections of thought and detail are inevitable in this
attempt.

Early history
Meinzer (1934) and Hall (1954) gave informative surveys
of hydrogeology prior to the 20th century. Until well into
the 16th century, it was believed that springs originated
from subterranean evaporation and condensation of sea
water, and that rain water was not sufficient to sustain the
observed flows. Bernard Palissy (15091589) of France
was the first to argue that springs were sustained by
rainfall. During the late 17th century, physicists Pierre
Perrault (16081680), and Edm Marriott (16201684)
pioneered quantitative hydrogeology. Based on rainfall
measurements, Perrault (1674) estimated total precipitation over the Seine river basin and concluded that rainfall
was adequate to maintain observed river discharge.
Marriott (1686) hypothesized that rainwater infiltrated
vertically down until it reached impermeable rock, then
moved laterally to replenish aquifers, springs and wells.
Hydrogeol J (2005) 13:724

In the early 19th century, the French pioneered welldrilling technology, encouraged by spectacular successes
with artesian wells in the Paris Basin. The French Science
Academy showed keen interest in data from the artesian
wells. One of the best studied wells at this time was at
Grenelle in the greater Paris area (Davis 1999). Dominique-Francois Jean Arago (17861853), noted physicist, studied temperatures of deep mines and water temperatures from artesian wells, and estimated the geothermal gradient in the Paris area to be 29.4 m/C (53.6 ft./
F). In 1835, he used Grenelle well data and calculated
that a 500-m deep well would provide water at a temperature that would be 18.8C above the ambient temperature at the land surface and well-suited for spaceheating during winter.
Alexander von Humboldt (17691859) of Germany,
who led a 5-year scientific expedition to South America,
observed in a cave in north-eastern Venezuela that the
water from a spring was colder than the ambient mean
annual temperature. He concluded that the spring water
was recharging at a higher elevation, and that the water
had not yet equilibrated with the host rocks (Davis 1999).
In Venezuela, he also measured temperatures of many hot
springs with near-boiling temperatures. Noting that the
nearest volcanic manifestations were far away, Humboldt
concluded that the high temperatures were due to deep
circulation of groundwater. Using European estimates of
geothermal gradient, Humboldt (1844) estimated the
source depth to be 2,176 m.
The modern conception of hydrogeological systems as
constituting the lower part of the hydrological cycle was
established by Marriott, Arago, and Humboldt. They
recognized deep infiltration of groundwater, and its return
to land surface because of permeability variations, and
geothermal heat.

Mathematization of hydrogeology
The 19th century witnessed the birth of quantitative hydrogeology. Fouriers (1822) heat conduction model was
extended to electricity by Ohm, to flow in capillaries by
Poiseuille and Hagen, and to molecular diffusion by Fick.
In this atmosphere, Henri Darcy (18031858), found that
the flux of water filtering through a sand column was
directly proportional to the gradient of hydraulic head,
now referred to as Darcys Law (Darcy 1856). Darcy was
the first to extend Fouriers law to flow of water in natural
porous materials, and to explicitly incorporate gravity in
defining hydraulic head. Much credit for bringing the
steady flow of groundwater within potential theory goes
to Jules-Juvenal Dupuit (18041866), who idealized a
permeable medium such as sand to be a collection of
small diameter tubes, and showed that Darcys Law was a
special case of de Pronys equation, with inertial effects
neglected (Dupuit 1857). In a work of great insight, Dupuit (1863) portrayed an artesian aquifer within the
framework of a groundwater basin, confined by clay
layers and connected to the water table at higher elevaDOI 10.1007/s10040-004-0422-5

9
Table 1 Some events of historical interest in hydrogeology

1580
1674
1686
1822
1828
1836
1844
1856
1857
1863
1877
1879
1896
1907
1907
1912
1922
1923
1923
1926
1928
1931
1933
1935
1940
1940
1941
1944
1946
1952
1953
1954
1954
1955
1956
1957
1959
1959
1960
1960
1960
1960
1960
1962

Hydrogeol J (2005) 13:724

Palissy argues that springs are fed by rain


Perrault estimates that rainfall is adequate to maintain stream flows in the Seine Basin
Marriott suggests that infiltrating rainwater moves laterally, sustains springs, aquifers,
and wells
Publication of Fouriers Analytic Theory of Heat
Du Commun explains the statics of freshwatersalt water interface in a coastal aquifer in
New Jersey
Arago provides estimate of geothermal gradient and discusses implications to artesian well
at Grenelle in the Paris area
Humboldt presents field evidence from Venezuela to assert that springs are manifestations
of deep groundwater circulation
Darcy describes the law relating water flux to hydraulic gradient in a porous medium
Dupuit idealizes sand as a complex of capillary tubes and describes Darcys Law as a special
case of de Pronys equation
Dupuit portrays an artesian basin and artesian wells in terms of hydraulic head profiles,
and solves the radial flow equation to steady flow in a aquifer with a free surface
Lucas defines hydrogeology
U.S. Geological Survey created on a recommendation by the National Academy of Sciences
Darton describes the regional flow system of the Dakota sandstone, including vertical leakage
through very low permeability shale layers
Invention of the deep-well turbine pump in California, inspired by irrigation demands
Buckingham defines a capillary potential for an unsaturated soil, and relates it to moisture
content and hydraulic conductivity
Lee prepares water budget for the Owens Valley groundwater basin, integrating surface water,
groundwater, evaporation, and consumptive needs of plants
Gardner invents the tensiometer to measure capillary potential, and presents moisture
characteristic curves measured for the first time
Terzaghi defines effective stress, and solves the one dimensional transient groundwater
flow equation to analyze consolidation of a clay column
Meinzers monograph on groundwater and its occurrence in the United States. Groundwater
hydrology takes formal form, as part of the hydrological cycle
First reported land subsidence at the Goose Creek oil field in Texas caused by subsurface fluid
(oil) production
Meinzers intuitive description of elastic deformation of aquifer material and expansion
of water during transient water production
Richards formally describes transient unsaturated flow as a non-linear partial differential
equation
First reported occurrence of land subsidence due to groundwater withdrawal in San Jose,
California
Theis publishes the non-equilibrium formula for non-steady radial flow to a well
Jacob gives physical meaning to Theis storage coefficient in terms of the vertical
compressibility of the porous medium, the compressibility of water, and the porosity
of the aquifer
Hubbert defines a fluid potential for groundwater systems, describes nature of hydraulic
conductivity, and suggests that the interface between freshwater and sea water will be inclined
under dynamic conditions of groundwater flow
Biot generalizes Terzaghis one-dimensional consolidation theory to three dimensions,
incorporating coupling between fluid flow and matrix deformation through effective
stress
Introduction of Piper diagram for interpretation of groundwater geochemistry
Jacob opens up groundwater hydraulics to leaky multiple aquifer systems
First successful attempt to solve the Richards equation for transient unsaturated flow
Hubbert extends theory of groundwater motion to oilwater contact in petroleum reservoirs
and significantly influences oil exploration
Scheidegger applies Brownian Motion idea to formulate the advectiondispersion equation
and defines the parameter dispersivity
Skempton defines the coefficients governing undrained response of water-saturated soils
Chebotarevs notion of metamorphism of natural waters along groundwater flow path
Bullard introduces heat flow probe for measurement of heat flux from ocean floor
Hubbert and Willis analyze hydraulic fracturing, and launch a new technique for measurement
of in situ rock stresses
Hubbert and Rubey propose a model for overthrust faulting drawing upon Terzaghis effective
stress
Hem and Cropper use the EhpH diagram for ferrousferric iron stability fields
Garrels monograph on mineral equilibria inspires a new generation of geochemical models
Clough introduces the finite element method
Bear introduces the dispersion tensor for macroscopic description of velocity-dependent
spreading of solutes in porous media
Barenblatt, and Warren and Root propose the double porosity model for naturally fractured
reservoirs
Back introduces the concept of hydrogeochemical facies in groundwater systems
Identification of interbasin groundwater movement in the Great Basin of Nevada
DOI 10.1007/s10040-004-0422-5

10
Table 1 (continued)

1962
1962
1963
1964
1966
1971
1976

Meyboom describes a prairee profile describing regional groundwater movement in


Saskatchewan, Canada
Toth describes the anatomy of regional groundwater flow systems
Failure of the Baldwin Hills reservoir near Los Angeles due to differential land subsidence
Tyson and Weber use computer models for groundwater basin evaluation
First report of earthquake swarms induced by deep-well injection at the Rocky Mountain
Arsenal well near Denver, Colorado
Failure of the Lower San Fernando Dam near Los Angeles due to soil liquefaction during
a major earthquake
First experimental demonstrating that earthquakes could be triggered by water injection

tions. Dupuit also pioneered mathematical analysis of


steady flow to wells in confined and unconfined aquifers.
The mathematical descriptions of Darcy and Dupuit
inspired investigations of seepage through soils, engineering geology, and water supply. Philipp Forchheimer
(18521933) of Austria perfected the application of flow
nets to analyze steady state seepage in engineered earth
systems. In Germany, Adolf Thiem and his son G nther
Thiem introduced the use of tracer tests and well tests to
design water supply systems.

Birth of hydrogeology in the United States


During early 19th century, natural springs and spas were
popular in the United States for their medicinal benefits.
The earliest chemical analyses of waters from springs
were reported by physicians as far back as 1827 (Davis
and Davis 1997). In California, the discovery of large
borax crystals in the muds of Borax Lake, and the discovery of travertine deposits by William Blake in the
Salton depression of the Imperial valley during the early
1850s were attributed to geochemical mixing of waters of
volcanic origin with waters of arid inland lakes. Another
insightful early American contribution was that of Du
Commun (1828). In 1824, an artesian well had been
drilled at New Brunswick, New Jersey by Americas first
well-driller Levi Disbrow (Carlston 1963). By 1827 observations from this 67-m (220-foot) well showed that the
water level rose from 2.4 to 4.3 m (814 feet) above the
surface of the local stream, and that the well discharge
rose and fell exactly and continually with the ocean tides.
Joseph Du Commun, a French instructor of the West
Point Military Academy, explained the observed phenomenon with the simple device of a U-tube in which two
immiscible liquids (saltwater and freshwater) were in
hydrostatic equilibrium. He extended the idea to the
coastal interface between freshwater and saltwater, and
presented quantitative estimates based upon the density
contrasts of saltwater and freshwater. Du Communs estimates preceded, by several decades, the commonly accepted contributions of Ghyben (1888) and Herzberg
(1901) on the static interface between freshwater and
saltwater.
Prior to 1870, science in the United States was largely
in the domain of the military. This situation changed in
1879 when the Congress created the U.S. Geological
Survey on a recommendation by the National Academy of
Hydrogeol J (2005) 13:724

Sciences. The nations boundaries were rapidly expanding


to unexplored territories in the west, and there was an
urgent need for classification of public lands, and the
examination of geological structure and mineral resources
of the national domain. The motivation was to educate
and assist the settlers in a new land towards a wise use of
natural resources. With the inception of the U.S. Geological Survey, significant amounts of public funds came
to be invested in the exploration of water resources for
public benefit. Ever since, the U.S. Geological Survey has
played a major role in the growth of hydrogeology in the
U.S. and in other parts of the world.
The early hydrogeological investigations were disposed towards recognizing hitherto unknown phenomena,
and to understand the role of groundwater within the
hydrological cycle. As early as 1888, Franklin H. King
had correlated diurnal water table fluctuations with discharge of water by vegetation (transpiration). King (1892)
had also discovered that in an observation well situated
near a railroad, the water level rose whenever a train went
by, but fell again as soon as the train had passed. An early
theoretician in groundwater hydrology was C.S. Slichter
(18641946), a mathematician who used complex variable theory to analyze two-dimensional steady state flow
systems (Slichter 1899).
Grove K. Gilbert (18431918), and Israel Russell
(18521906) ushered in paleohydrology with their studies
in the Great Basin. Gilbert (1890) deciphered the quaternary history of Lake Bonneville in Utah, shaped by a
combination of glaciation, tectonics, and erosion. Inspired
by Gilberts work, Russell (1885) recognized a similar
history of ancient Lake Lahontan in northwestern Nevada
during the Quaternary.
A bulk of the systematic hydrogeological work
throughout the nation was carried out by a band of USGS
geologists. Notable among these were Nelson Darton
(18651948), Walter Mendenhall (18711957), and
Charles Lee (18831967). Darton (1896), and coworkers
deciphered the artesian aquifer system of the Dakota
Sandstone, recharged over the Black Hills on the west,
and having areas of discharge 300500 km to the east in
South Dakota (Bredehoeft et al. 1982). Darton intuitively
explained the existing regional hydraulic gradient to the
east, and upward leakage through a few hundred meters of
overlying clay formations. Mendenhall, who mapped
groundwater systems in the San Joaquin Valley and in the
San Bernadino Valley of California recognized that the
aquifer systems of these valleys, directly connected to
DOI 10.1007/s10040-004-0422-5

11

surface water, were not only large storage reservoirs, but


were also well suited for regulating flow to aid irrigation
and other needs. Between 1908 and 1913, when the Los
Angeles Aqueduct was built to bring Owens River water
to Los Angeles, Lee (1912) carried out a detailed water
budget of the groundwater basin, considering surface
water flows, evaporation, transpiration by plants, and
groundwater recharge and discharge.
The most noteworthy among these hydrogeologists
was Oscar Meinzer (18761948), who had the vision and
the scholarship to appreciate groundwater in its totality,
including its connections to surface water and the plant
kingdom. His elucidation of elastic deformation of aquifers, and its role in the transient response of aquifers are
indicative of a field geologist with admirable intuition.
His 1923 publication on the general principles of
groundwater occurrence and movement, and his 1927
publication on plants as groundwater indicators are remarkably relevant today as our awareness of ecosystems
increases.

Soil physics and the vadose zone


By late 19th century, the U.S. Department of Agriculture
was investing substantial resources into improving irrigated agriculture through scientific research. At that time,
agricultural scientists were puzzled by an observational
paradox: soils of arid regions, at depths a little below the
surface, were found to be generally wetter and held their
moisture for much longer periods than soils of humid
areas in the dry season. Edgar Buckingham (18671940)
of the Bureau of Soils, a physicist, helped clear this
paradox. Applying the concept of capillarity, and using
Fouriers model of thermal conduction, Buckingham
(1907) defined a capillary water potential for a soil, and
demonstrated that water movement in soils is driven by a
combination of gravity and the gradient of capillary potential, and that soil hydraulic conductivity was a function
of capillary potential.
However, Buckingham did not show how capillary
potential could be measured. He simply inferred it from
capillary rise in hydrostatic columns. Willard Gardner
(18831964) overcame this shortcoming by inventing the
porous-cup capillary potentiometer, now known as the
tensiometer (Gardner et al. 1922). For the first time, it
became possible to quantitatively express soil moisture
content and hydraulic conductivity in terms of a measurable quantity, the capillary potential. A decade later,
Richards (1931) formally wrote down the non-linear
partial differential equation for three-dimensional moisture movement in soils.

Transient groundwater flow


It became obvious in the early 20th century that the flow
of water in hydrogeological systems was time-dependent,
and that the steady-state assumption of potential theory
Hydrogeol J (2005) 13:724

was too limiting. Handling such transient systems required a new hydraulic property, analogous to specific
heat in the heat equation. The identification of such a
property, the hydraulic capacitance, evolved over three
decades from parallel developments soil mechanics, soil
physics, and groundwater hydrology (Narasimhan 1986).
The first person to successfully apply the transient heat
equation to fluid flow in a porous medium was Karl
Terzaghi (18831963), who solved the one-dimensional
clay consolidation problem of water expulsion caused by
an imposed external load. In this work, Terzaghi (1923)
elucidated analogies among temperature and pore pressure, heat content and water content, thermal conductivity
and clay permeability, and specific heat and clay compressibility.
In the artesian basin of South Dakota, Meinzer (1928)
found that the time taken for stabilizing artesian heads
after opening or capping a well was longer in the case of
relatively more compressible fine-grained formations than
in more stiff coarse-grained materials. He attributed this
to the differences in the ability of the different formations
to take water into storage. He also reasoned that the
ability to take water into storage involved a combination
of change in porosity and expansion of water. Soon
thereafter, Charles Theis (19001987), treated change in
storage due to change in hydraulic head as analogous to
specific heat, and applied the heat equation to transient
flow of water to a well. In this work, Theis (1935) was
helped by mathematician Clarence Lubin of the University of Cincinnati (Freeze 1985). In the illustrative application, Theis used data from a water-table aquifer, and
estimated its specific yield. Jacob (1940) provided a rigorous interpretation of Theis specific storage coefficient
by combining vertical compressibility of the porous medium and water compressibility. Jacobs analysis showed
that Theis formulation was appropriate for confined
aquifers, rather than water table aquifers. Theis launched
groundwater hydrology in the new direction of well hydraulics, which dominated hydrogeology over the next
half a century. Jacob (1946) and Mahdi Hantush (1921
1984) played major roles in this by extending transient
flow analysis to multiple aquifer systems.
Transient flow in unconfined aquifers demanded a
more involved mathematical analysis because of timedependent drainage from the unsaturated zone. To keep
mathematics tractable, Boulton (1954) assumed that
drainage from above the water table was exponentially
related to time, leading to a mathematical form in which
hydraulic capacitance (specific yield) amounted to a sum
of an instantaneous constant value and an exponential
dependence on time. Among other developments in wellhydraulics, mention may be made of the transient analysis
of water level fluctuations in response to seismic waves in
a well piercing confined aquifer by Cooper et al (1965).
This analysis showed that under certain conditions, seismic signals can be amplified by wells. The first attempts
to solve the non-linear partial differential equation for
transient flow in unsaturated soils were made by Klute

DOI 10.1007/s10040-004-0422-5

12

(1952) and Philip (1956), who pioneered modern theory


of infiltration in soils.
By the 1970s, the transient equation was being used in
two forms, one for fully saturated media, and the other for
unsaturated media. In the former, the hydraulic capacitance (specific storage) included deformation properties
of the porous matrix and water. In the latter, hydraulic
capacitance was entirely accounted for by the rate of
change of moisture content. A single transient flow
equation uniting the saturated and unsaturated regions
was lacking because of the differences in the storage
mechanisms of the two regimes. Drawing upon Bishop
and Blights (1963) extension of Terzaghis effective
stress principle to unsaturated materials, Narasimhan
(1975) defined a generalized hydraulic capacitance which
accounted for compressibilities of the porous medium and
water, and desaturation of the soil, integrating transient
flow in the saturated and the unsaturated domains.

Deformation of water-saturated porous media


From about the 1880s, civil engineers were studying
ground settlement in relation to embankments, dams, and
other structures built on water-saturated clays. However, a
credible quantitative theory was lacking to explain the
deformation of water-saturated geological formations.
During the 1920s, Terzaghi revolutionized the study of
soil deformation through insightful experiments on watersaturated clays, and founded the field of soil mechanics.
Central to his contribution was the hypothesis that volumetric strain in a porous medium was a function of the
difference between the external stresses acting on the soil
and the water pressure, defined as the effective stress
(Terzaghi 1923). Going beyond volume change, Terzaghi
also established that the frictional strength of water-filled
shear planes in soil masses was dependent on effective
stress acting on the plane. This finding soon became the
basis for explaining landslides, earthquakes, and the formation of folded mountains such as the Alps and the
Himalayas. Maurice Biot (19051985) extended Terzaghis one-dimensional theory to general three dimensional
deformation., accounting for the response of a watersaturated soil to loads imposed on the boundaries and the
three dimensional stressstrain responses associated with
the drainage of water (Biot 1941).
A water-saturated soil subjected to a change in external
stress (e.g., ocean tides, passing trains, earth tides), responds in such a way that part, if not all, of the imposed
stress is instantaneously borne by the water, resulting in a
step-wise change in water pressure. Subsequently, the
excess stress is gradually transferred to the solid skeleton
as the water drains with time, until the excess water
pressure is bled off. An analysis of undrained loading,
combining the effects of shear stresses and normal
stresses was presented by Skempton (1954).
Terzaghis theory, embellished by the contributions of
Biot, Skempton, and Bishop, largely forms the modern

Hydrogeol J (2005) 13:724

basis for quantitatively analyzing passive and catastrophic


deformation of hydrogeological systems.
The 1950s saw two papers spearheaded by Marion
King Hubbert (19031989). About this time, oil-well
drillers had discovered that reservoir rocks could be
fractured by injecting water under high pressure, and that
the artificial fractures enhanced reservoir productivity. By
analyzing stress distribution along the perimeter of the
well-bore during water injection, Hubbert and Willis
(1957) showed that a system of two vertical fractures will
initiate and propagate perpendicular to the direction of the
least horizontal principal stress. In areas of regional
compression where the least of the three principal stresses
is vertical, the vertical fracture would rotate and become
horizontal. In addition to placing the technology of reservoir stimulation on rational foundations, the Hubbert
Willis contribution opened up a new way of measuring in
situ tectonic stresses in rocks.
The hydraulic fracturing work led Hubbert to use
Terzaghis effective stress as a means of logically explaining the puzzling paradox of the transport of huge
blocks of the earths crust over long distances along low
angle thrust faults in orogenic belts. Hubbert and Rubey
(1959) hypothesized that the frictional strength of fault
planes was directly related to effective stress acting normal to the fault plane, and that if fluid pressures within
the plane became sufficiently high, the strength could
dramatically decrease, enabling the transport of enormous
rock blocks over long distances. Today, the effective
stress concept, in conjunction with MohrCoulomb failure criterion constitutes the standard model to explain
such phenomena as landslides, natural and triggered
earthquakes, and orogeny.
The turn of the 20th century witnessed an explosion in
oil production in many parts of the United States. In
parallel, the growth of irrigated agriculture in California
spawned the invention, around 1907, of the deep-well
turbine pump (Freeman 1968), which enabled water to be
lifted from great depths. Within a decade, the effects of
oil production and groundwater overdraft led to noticeable environmental effects. The first observed land subsidence due to subsurface fluid withdrawal was reported
by Pratt and Johnson (1926) from the Goose Creek oil
field, located at the head of Galveston Bay, Texas. In
addition to a 1-m deep subsidence bowl, earth-fissures,
presumably caused by differential subsidence, were also
observed. The first recorded land subsidence due to
groundwater withdrawal was fortuitously discovered
around San Jose, California in 1932 by the U.S. Coast and
Geodetic Survey, during repeat precision-leveling (Rappleye 1933). Meinzer (1937), based on his experience in
South Dakota, explained the observed more than 1.2 m
(4 feet) of land subsidence as due to the compaction of
fine-grained sediments, caused by declines in fluid pressure associated with groundwater overdraft. A catastrophic consequence of land subsidence was the failure of
the Baldwin Hills reservoir near Los Angeles in 1963.
Excessive pumping of oil had led to differences in the
magnitude of land subsidence on either side of a normal
DOI 10.1007/s10040-004-0422-5

13

fault, which triggered vertical displacement, leading to


failure of the dam.
Evans (1966) reported, for the first time, a strong
correlation between deep-well injection of water in the
Rocky Mountain Arsenal Well near Denver, Colorado,
and a swarm of over 700 earthquakes during a 2-year
period. Following Hubbert and Rubey (1959), the quakes
were attributed to gravity sliding along relatively steep
faults whose frictional strength had been decreased by
increase in water pressure caused by injection. The
credibility of the TerzaghiHubbertRubey hypothesis
was experimentally established by controlled fluid injection in an oil field at Rangely, Colorado by the U.S.
Geological Survey (Raleigh et al. 1976). These ideas were
later extended to understand earthquakes triggered by
man-made reservoirs beneath which large water pressures
may be generated either due to undrained response, or due
to transient groundwater motion (Simpson and Narasimhan 1990).
An interesting manifestation of undrained response is
the subtle, but measurable systematic response of fluid
pressure in confined aquifers due to variations in earth
tides, caused by the gravitational pull of the sun and the
moon. Making the simplifying assumption that stresses
associated with earth tides act horizontally, and that the
earth is free to deform in the vertical direction, Bredehoeft (1967) provided a method of estimating the storage
coefficient of the aquifer from the response of a well to
earth tides.
A more dramatic manifestation of response to undrained loading of a water-saturated granular soil came to
light with the failure of the Lower San Fernando Valley
dam near Los Angeles following an earthquake on
February 9, 1971. This hydraulic-fill earthen dam failed
due to the liquefaction of coarse granular material on the
upstream side of the dam, caused by excessive pore
pressures generated by the shaking motion and the consequent densification of the soil without adequate time for
water to drain (Seed et al. 1973).

Regional groundwater motion


In parallel with developments in well-hydraulics, Hubbert
(1940) invigorated interest in looking at groundwater
systems as a whole. He interpreted groundwater motion
dynamically in terms of impelling and resistive forces,
and an energy potential, with the systems bounded by
groundwater divides and impermeable barriers. Within,
isopotential surfaces and flow lines that refracted at material interfaces characterized the flow pattern. Analyzing
force balance at the immiscible interface between freshwater and saltwater, Hubbert showed that it will be inclined under conditions of groundwater flow. Hubbert
(1953) extended the freshwatersaltwater interface idea to
the oilwater interface in petroleum reservoirs, and elucidated the segregation, migration, and entrapment of oil
and gas under the influence of regional groundwater
motion.
Hydrogeol J (2005) 13:724

Meanwhile, geochemists studying groundwater systems began recognizing connections between variations in
groundwater quality and groundwater movement. Chebotarev (1955) noted that the anionic content of groundwater revealed much about the chemical processes to
which the moving groundwater is subject. Based upon a
large number of observations, he suggested that groundwater starts off by being rich in bicarbonate in areas of
recharge, and successively acquires sulfates, followed by
chloride. Thus, at discharge locations far removed from
recharge, water is enriched in chloride. Back (1960),
based on the disposition of chemical analysis data showed
that the outcrops of Cretaceous and Eocene sediments in
the Atlantic Coastal Plain constitute areas of discharge.
He suggested that the outcrop area of an artesian aquifer
need not necessarily be an area of recharge.
Following the detonation of underground nuclear devices in 1957, the U.S. Geological Survey began systematic groundwater studies of the Nevada Test site and
its vicinity, with the objective of assessing the impacts of
the tests on local groundwater resources. Aided by water
level data from shallow and deep wells, and chemical and
isotopic data, these studies soon led to the identification
of intra basin and interbasin regional flow systems in this
arid region, over several intermontane valleys extending
over thousands of square kilometers. Winograd (1962)
presented some of the early evidence of deep interbasin
flows.
These ideas came together in the Hydrology Symposium held at the University of Alberta, Canada, late in
1962. Toth (1962) presented results of hypothetical steady-state calculations for a rectangular, homogeneous flow
domain subject to sinusoidal boundary conditions imitating a fluctuating water table. By varying the dimensions of the domain and the parameters of the sinusoidal
wave, he showed that groundwater flow patterns can be
organized into shallow, intermediate and deep flow systems, and that such a recognition can help in deciphering
the relationships among groundwater circulation, water
quality, distribution of mineral phases, and spatial variations of plant communities. In a companion paper Meyboom (1962), presented field data to illustrate regional
groundwater flow patterns associated with a prairie profile in Saskatchewan, governed by the extensive presence
of a permeable horizon at depth, connecting the area of
recharge with an area of discharge located far away. This
work used data from nested piezometers, designed to
discriminate vertical components of flow. The prairie
profile suggested the possibility of groundwater flow
systems on a continental scale. Using the digital computer, Freeze (1966) illustrated the practical utility of the
concept through extensive parametric studies on two-dimensional steady-state systems.

Some examples
Abnormally high fluid pressures in some sedimentary
basins constitute unusual hydrogeological environments.
Among the possible causes of such systems, one is the
rapid build-up of sediment over-burden which causes
DOI 10.1007/s10040-004-0422-5

14

undrained pore pressure generation that may dissipate


gradually over geological time. These manifestations
suggest that large hydrogeological systems need to be
treated as transient on appropriate time scales. Toth and
Millar (1983) suggested, based on analysis of pore pressure data from depths in excess of about 1,220 m
(4,000 feet) in the Red Earth Region of Alberta, Canada,
that the observed fluid pressures are still responding to
erosional changes to topography initiated 112 million
years before present.
During the early 1970s, the Arab oil embargo motivated serious study of geothermal systems in the United
States. Geothermal systems fall into two categories: liquid-dominated and vapor-dominated. In the Imperial
Valley of California, which is part of an incipient
spreading center covered by rapid sedimentation by the
Colorado River, hydrothermal convective systems with
temperatures of 340C (640F) are known at depths of
2,1304,570 m (about 7,00015,000 feet). Exploratory
drilling during the 1970s revealed that the hot brines, 8
10 times as saline as sea water, and charged with metals
such as gold, silver, copper, and zinc were manifestations
of active ore-forming fluids.
More remarkable than the liquid-dominated systems is
the vapor-dominated geothermal system at the Geysers,
some 110 km (about 70 miles) north-east of San Francisco. The largest of its kind in the world, the Geysers
represents an abnormally under-pressured natural system.
Here, no water table has been encountered down to more
than 3,660 m (about 12,000 feet). Before active production, steam pressure and temperature were, respectively,
243C (470F), and about 3.45 mPa (500 pounds per
square inch), even in wells over 2,440 m (8,000 feet)
deep. In contrast, the expected hydrostatic pressure at
2,440 m (about 8,000 feet) would be in excess of
20.7 mPa (3,000 pounds per square inch). The geological
conditions that give rise to this type of under-pressuring is
as yet not understood. Based on mathematical modeling,
Pruess and Narasimhan (1982) postulated that vapor
production from wells was the consequence of heat and
mass transport in a double-porosity fractured porous
medium.
The Mississippi Valley-type massive, stratiform lead
zinc deposits found distributed around the peripheries of
many sedimentary basins (e.g., Michigan Basin, Illinois
Basin, Anadarko Basin) in central North America constitute hydrogeological systems of continental proportions. They are epigenetic deposits, formed at temperatures of 100150C, at relatively shallow depths due to
the action of chloride rich brines, presumed to have
originated at great depths within sedimentary basins. A
key question concerns the mechanisms by which the
brines are transported on a continental scale from depths
in the center of the basin to the near-surface around the
periphery to cause ore deposition. Cathles and Smith
(1983) suggested, based on computer models, that these
deposits formed due to episodic expulsion of ore-forming
fluids, separated by long quiescent periods of fluid pressure build-up and rupture.
Hydrogeol J (2005) 13:724

Geochemistry of hydrogeological systems:


aqueoussolid chemical interactions
A discussion of 20th century geochemical developments
in hydrogeology can be found in Back and Freeze (1983).
An early interpretive view of groundwater chemical
analysis was made by Palmer (1911), who postulated that
salts dissolved in natural waters constitute a chemical
system of balanced values. The next four decades witnessed a recognition that water quality data indicate
predictable interactions among rain water, soil, and the
rock types in the path of groundwater. Rogers (1917)
compared the chemistry of shallow waters of the San
Joaquin Valley, California with deeper oil field waters
and concluded that due to the reducing action of oil and
gas, sulfate, a dominant anion of the shallow waters, had
been reduced to hydrogen sulfide. He also speculated on
the role of sulfate reducing bacteria. Renick (1924)
studied groundwater in Tertiary formations of east-central
Montana and found that shallow waters, rich in calcium
and magnesium, progressively became enriched with
depth in sodium at the expense of calcium and magnesium due to base ion exchange facilitated by the presence
of certain silicate minerals. These observations motivated
the development of systematic procedures for the classification and graphical presentation of chemical data to
facilitate interpretation. A pioneering work is that of Piper
(1944), who showed how tri-linear diagrams could be
used to screen a large number of water-analyses for
critical study with respect to sources of the dissolved
constituents, modifications in chemical character as water
passes through an area, and related geochemical problems.
During the late 1940s geochemistry became established as a separate discipline, and rapidly asserted its
influence in the earth sciences. Accumulating data from
oil-field brines, and associated zones of anomalously high
water pressures stimulated considerable interest. Berry
and Hanshaw (1960) studied abnormally high pore pressures from three widely separated areas in North America
and argued that pressure-salinity anomalies in deep sedimentary basins can be accounted for by compacted shale
acting as selectively permeable membranes. White (1965)
surveyed brines from several sedimentary basins in North
America, and concluded that the early diagenetic changes
in these waters were significantly influenced by organic
content and bacterial activity, as well as ion exchange. He
suggested that many characteristics of saline waters are
best explained by the hypothesis that fine-grained sediments act as semi-permeable membranes permitting selective escape of water, carbon dioxide, sodium, boron
and sulfur, relative to chloride and calcium.
John Hem (19161994) made two influential contributions to groundwater chemistry. In the first, Hem
(1959) summarized the knowledge and experience of the
U.S. Geological Survey in sampling, analyzing and interpreting groundwater quality. In the second, Hem and
Cropper (1959), introduced the use of EhpH diagrams to
interpret the aqueoussolid stability relations of redox
DOI 10.1007/s10040-004-0422-5

15

species in groundwater systems. Contemporaneously,


Garrels (1960) introduced the Eh-pH diagram based on
thermodynamic considerations, and demonstrated its
power in understanding diagenesis, low-temperature ore
deposits, and the genesis of soils. Garrels work soon led
to rapid developments in computer-based mathematical
models. Helgeson (1968) was among the earliest to
translate the thermodynamic equilibrium interactions into
a set of partial differential equations amenable to be
programmed on a digital computer. Geochemical analysis
of aqueoussolid interactions require consideration of
other processes as well. These include adsorption on
mineral surfaces, and ion-exchange between water and
minerals such as zeolites.
Micro-organisms play important roles in earth processes as widely varying as weathering and soil genesis,
formation of rocks and minerals, and genesis of oil deposits. It is now believed that they thrive in aqueous environments at depths of 3 km or more, and at temperatures
of more than 66C (150F). Although observational
knowledge of these organisms is accumulating, little is
known about the way they function. Microorganisms
extract energy for their sustenance from specific chemical
reactions that would otherwise release energy slowly.
The inventions of the atomic mass spectrometer and
the electron microprobe during the 1940s opened up new
ways of understanding the functioning of hydrogeological
systems. The former helped precise measurement of stable and unstable isotope abundances, while the latter
helped high resolution profiling of chemical concentrations within mineral grains. Over the past six decades,
these instruments have seen spectacular improvements.
Besides providing clues about the age and source of the
waters, these instruments have helped greatly in understanding the evolution of the liquid phase as well as the
mineral phases in hydrogeological systems.
It is fitting to conclude a discussion of geochemistry
with the grandest of all hydrogeological problems, the
origin of oceanic and continental crusts. Deming (2002)
pointed out that plate tectonics was established in the
Earths crust only some 2.5 billion years ago, at the beginning of the Proterozoic, and that the nature of the crust
during the previous 2 billion years of the Earths history is
largely unknown. The continental crust is believed to
have come into existence through hydrolysis of the
primitive crust. The large volumes of water required, it is
speculated, came about slowly from extra-terrestrial accretion.

Chemical transport
Within a hydrogeological system, aqueoussolid interactions occur at a given location. For these reactions to
influence the regional system, the reaction products have
to be transported. Among groundwater hydrologists, the
impetus to consider chemical transport emerged during
the 1950s with concerns about water quality problems of
saltwater intrusion. Migration of dissolved substances in
groundwater involve a combination of bulk movement
with the flowing water at average water velocity, referred
Hydrogeol J (2005) 13:724

to as advection, and concomitant diffusive spreading.


Scheidegger (1954) looked at Darcys Law from the
perspective of Brownian Motion, and by superposing
random motion of particles over average bulk motion
(referred to as drift) showed that the random motion locally introduces a dispersion, which can be quantified by
a new parameter, dispersivity. For homogeneous media,
dispersivity can become a constant under some statistical
assumptions. This idea was extended to describe the migration and spreading of dissolved contaminants in geological materials by Bear (1960), who showed that velocity controlled spreading in a homogeneous isotropic
groundwater body can be represented with the help of a
second-rank dispersion tensor with longitudinal and
transverse components. The sum of this dispersion and
molecular diffusion is referred to as hydrodynamic dispersion. The velocity-dependent dispersion coefficient
becomes a strong function of scale in heterogeneous
media. To gain insights into scale-dependence, Gelhar et
al. (1979) applied stochastic calculus to the steady flow of
water in a stratified aquifer and found that for large time
periods, the dispersion coefficient approaches a constant
value that depends on the statistical properties of the
medium.
Contamination of shallow and deep groundwater systems occur due to a variety of human activities including
mining, agriculture, landfills, disposal of toxic wastes,
septic tanks and animal wastes, and industrial effluents.
Abandoned mines, mine wastes, and solid wastes from
thermal power plants generate highly acidic effluents that
transport toxic elements such as arsenic, boron, lead,
chromium, copper, molybdenum, zinc, and so on. During
the 1980s, contamination of groundwater by organic
chemicals came to be recognized as a very serious environmental concern. The sources of these contaminants
include leaky gasoline tanks, organic solvents and other
effluents from manufacturing plants, and electrical
transformer fluids. Some of these chemicals are known to
be carcenogenous at concentrations as low as a few parts
per billion. Although most of these organic compounds
will ultimately decay to carbon dioxide or methane, very
little quantitative data are available about the decay paths
of most of these contaminants. Compounding the problem
is the fact that new chemicals are introduced into the
human habitat by the tens of thousands each year.

Discontinuous media
Geological formations with fractures, joints, solution
cavities, and large openings formed due to various causes
may be collectively referred to as discontinuous media.
The nature of occurrence of water in these media, and
exploitation of water residing in them continue to intrigue
and challenge hydrogeologists. Unlike sediments such as
clay or sand that can be visualized as continuous or
homogeneous in small samples, discontinuous media
defy such visualization even in very large volumes. Thus,
these systems are not amenable to mathematical analysis
DOI 10.1007/s10040-004-0422-5

16

with methods that have otherwise proved to be successful


in analyzing physical systems. The discontinuous nature
also poses challenges to exploration for water.
Discontinuous media may be divided into two groups.
Karst and Karst-like systems are understood on a scale
ranging from hundreds of meters to kilometers. Fractured
media and fractured porous media are understood on a
scale of centimeters to tens of meters. Karst-like landscapes are characterized by closed depressions integrated
with underground drainage, with disappearing surface
streams and caves (White 1988, p. 347). The large underground drainage channels may have originated through
solution of carbonate rocks (true Karst), solution of
evaporite deposits and salt domes, or may have formed
due to rapid cooling of thick lava flows (lava tubes).
Whereas fractured media do not generally constitute
highly productive aquifers, Karst-like formations may
form highly productive aquifers and may sustain large
springs. In Nevada and California, highly productive,
deep carbonate aquifers are known to transfer groundwater across several water divides (Eakin 1966).
Somewhat more amenable to mathematical analysis
are systems in which consolidated, brittle rock formations
have been mechanically broken by fractures or joints that
usually occur in sets. Members of each set have similar
attributes of orientation and size. Despite this broad regularity, individual fractures pinch out along their length
and breadth, and their apertures vary spatially. These
pinch-outs and aperture changes are difficult to characterize and predict, and herein lie the hydrogeological attributes peculiar to fractured rock systems. The fracture
porosity of rocks is usually less than one percent, and
because the compressibility of fractures is very small (of
the same order as compressibility of water), the hydraulic
capacitance of these rocks is very small. However, individual fractures can transmit water with great ease.
Hence, fractured rocks tend to possess relatively high
hydraulic conductivity. The high hydraulic conductivity
and the low porosity and compressibility of the rocks
combine to bestow these rocks with very high hydraulic
diffusivity. Consequently pressure transients can migrate
very rapidly through fractured rocks.
The bulk hydraulic conductivity of fractured rocks
stem from a combination of fracture apertures, fracture
lengths, their number and connectivity. Idealizing a
fracture as a special case of a water-filled rectangular
tube, Boussinesq (1868) showed that the flux through a
fracture is proportional to the third power of its aperture.
This observation is often referred to as the cubic law.
Despite the difficulties inherent in assigning a single aperture magnitude to a rough-walled fracture, the cubic law
is still used as an approximation. Because fractures can
conduct contaminants rapidly, the study of fractured rocks
has attracted much attention since the 1960s. Snow (1965)
revived interest in the parallel plate model in applying it
to problems of engineering geology and groundwater
hydrology. Snow (1969) also showed that an idealized
fracture set, with continuous fractures of uniform aperture
can be effectively treated as an anisotropic medium. At
Hydrogeol J (2005) 13:724

present, statistical methods are widely used to understand


the connections between statistically described fracture
networks and scale-dependent hydraulic conductivity of
bulk rock.
Materials such as sandstone or shale that have primary
porosity, when fractured, give rise to fractured porous
media. In such media, a bulk of the stored fluid (water,
oil, or gas) resides in the porous matrix, but the transmission of the fluids over long distances takes place
through the interconnected fractures. The low-diffusivity
matrix blocks occur as islands in the high diffusivity
fracture network. Barenblatt et al. (1960) and Warren and
Root (1963) proposed the double-porosity model to describe transient fluid flow in such media, which helped
preserve the effects of fine-scale structure on large scale
pressure transients.
The hydraulic attributes of a fracture dramatically
change when it changes from a state of full water saturation to one of partial saturation. In this case, fractures
become strong hydraulic resistors in comparison with the
rock blocks. Under such conditions, water tends to move
very slowly through the rock blocks and along films on
the walls of the fractures. This remarkable attribute has
played a role in the selection of the Yucca Mountain site
in Nevada by the U.S. Congress in 1987 as the most
desirable site to be investigated for the disposal of high
level commercial radioactive wastes. At Yucca Mountain,
the water table lies at a depth of 457 m (about 1,500 feet),
in a desert area with very low rainfall. The mountain itself
comprises extremely fine-grained fractured ash-fall tuffs.
Both the unsaturated fractures and the porous matrix are
characterized by extremely low hydraulic conductivities.
Additionally, one of the ash-fall layers below the proposed repository depth is rich in zeolites, capable of adsorbing significant quantities of toxic cations. The problem of radioactive waste disposal at Yucca Mountain as
well as at other places within and outside the U.S. has
fostered substantial research to understand the movement
of water and contaminants in fractured rock systems. The
movement of water and contaminants in these systems
continues to elude precise mathematical description.
Nonetheless, the licensing process for authorizing the
commissioning of the site requires assurance that toxic
contaminants will not reach the accessible environment
for 10,000 years or more. On this time scale, the difficulties of fracture flow characterization are compounded
by uncertainties of future climatic changes.

Marine hydrogeology
By the end of the World War II, geologists came to recognize that the large scale structure of the Earth was much
simpler beneath the oceans than on the continents. This
motivated active scientific exploration of the ocean bottom during the 1950s. The exploration of the ocean bottom is so intimately tied with understanding fluid circulation in the oceanic crust that a significant part of marine
geology and marine geophysics comes within the scope of
DOI 10.1007/s10040-004-0422-5

17

hydrogeology. A milestone in the history of marine geophysics was the invention of a heat-flow probe by Edward
Bullard (19071980), which enabled the measurement of
natural heat flux from the oceanic crust (Bullard et al.
1956). During the 1960s, measurements made with this
probe and its variants from the Atlantic, the Pacific, and
the Indian oceans helped greatly in the establishment of
the theory of plate tectonics.
Modern exploration of the ocean bottom includes
drilling of boreholes into the oceanic crust, supported by
detailed studies of the ocean floor in the vicinity of the
borehole. The boreholes are subjected to detailed geophysical and geochemical logging, and hydraulic packer
tests where possible, to estimate in situ hydraulic conductivity. Frequently, detailed heat flow measurements
are made in the vicinity of the boreholes to prepare heat
flow contour maps. These data constitute the infrastructure to understand circulation of water in the oceanic
crust.
Marine hydrogeological systems can be divided into
those that are associated with crustal extension (with
plates moving away from a rift), and those associated with
regions of subduction. In regions of crustal extension, a
principal goal is to understand the attributes of the convective cells, as one moves progressively away from the
ridge axis. The zones of subduction are regions where
meteoric water begins its migration towards the mantle.
One interesting hydrogeological feature of subduction is
that the gradual displacements along the plane of underthrusting, in conjunction with the very low permeability
of the sediments, can lead to the generation of abnormally
high pore pressures, leading to the occurrence of mud
volcanoes under the weight of several thousand meters of
sea water.
Because of the difficulty of accessing the ocean bottom, marine hydrogeology has inspired the invention of
novel tools and technologies. As an example, one may
cite certain special boreholes of the Ocean Drilling Project. Drilled below water depths in excess of 3,660 m
(about 12,000 feet), these boreholes can be periodically
reentered after intervals of many months to years to
deepen the borehole, run repeat logs, or carry out various
types of tests. Hydrogeologists used to studying continental systems, have much fascinating knowledge to gain
from developments in marine hydrogeology.

Numerical methods
The mathematical techniques which Fourier pioneered to
solve the diffusion equation are practically useful in those
hydrogeological systems with known symmetry and heterogeneity. The solution techniques are inadequate for
regional systems characterized by arbitrary geometry,
heterogeneity, and boundary conditions. Prior to the
1970s, efforts were made to overcome this deficiency
with the help of physical models and analog models.
However, the rapid growth of digital computers starting
from the late 1960s helped establish numerical methods as
Hydrogeol J (2005) 13:724

the most powerful approach for modeling hydrogeological systems. The following discussion is restricted to
numerical methods.
The use of the digital computer to solve the transient
diffusion equation was pioneered during the World War II
by von Neumann using the finite difference approach, in
which spatial gradients are approximated along principal
directions by finite differences. However, it was too
limiting to restrict grid points along the principal directions of the coordinate axes, especially when the flow
domain had a complex shape. This limitation was overcome by taking a physical, intuitive view of the governing
equation as a mass conservation statement over an elemental volume of known geometry, whose surface is divided into a finite number of segments. The earliest
proponent of this approach was MacNeal (1953). This
intuitive approach was first applied to groundwater basins
by Tyson and Weber (1964), who used polygonal volume
elements. The shortcoming of the method is that it cannot
conveniently evaluate fluxes in anisotropic media. The
finite element method, developed by Clough (1960) for
solving stressstrain problems in solid structures, provided a convenient way of handling anisotropic media.
The central theme of this novel approach was to use a
small triangular region defined by three non-collinear
points as the basis of evaluating spatial gradients in any
direction, rather than using two points to evaluate a gradient in a fixed direction. This method was extended intuitively to solve problems of heat conduction in mass
concrete structures by Wilson (1968). Since the 1960s, the
Finite Difference approach, the approach of using
polygonal elements, and the finite element approach have
been employed to solve for steady and non-steady flow in
multidimensional systems involving saturated and unsaturated groundwater flow.
In general, hydrogeological systems are idealized as
interactions among: flow of multiple fluid phases, deformation of the porous medium, transport of heat, transport
of dissolved and suspended chemical compounds, and
aqueoussolid chemical interactions. Mathematically,
these interactions are represented by a set of partial differential equations, with each process being represented
by a diffusion-type equation, supplemented by advection
where appropriate. Chemical interactions are commonly
handled through equilibrium thermodynamic considerations, which include the Law of Mass Action, mass conservation, electrical neutrality, electron conservation, and
Gibbs phase rule. Additionally, surface processes such as
adsorption and ion exchange are handled through empirical relationships obtained from batch experiments. Departure from equilibrium are handled with kinetic coefficients. The interactions referred to above are quite
complex, and the various parameters relevant to the processes are not available a priori. Even now, most models
address only a few of these couplings at a time, as for
example, flow and deformation, multi-phase flow and
heat, and fluid flow with heat flow and chemical reactions.

DOI 10.1007/s10040-004-0422-5

18

The standard mode of application the numerical model


is to assume that the geometry, material properties, initial
conditions, and boundary conditions are known, and to
solve for the dependent variable, the hydraulic head. This
is the forward problem. In general, this problem is
mathematically well-posed, and unambiguous solutions
can invariably be found. However, one is often confronted
with the situation where the hydraulic conductivity and
the hydraulic capacitance are not a priori known. In this
situation, one seeks to use field measurement of the hydraulic head at a finite number of points to estimate the
magnitude of the hydraulic parameters, with the numerical model. This is referred to as the inverse problem. The
difficulty with the inverse problem is that its solution
inherently non-unique, even if data on hydraulic head,
system geometry, and boundary conditions are all known
in detail. Among the earliest to examine a groundwater
basin from the perspective of the inverse were Vemuri
and Karplus (1969). They considered transient flow in an
unconfined aquifer in the Eastern portion of the San
Fernando Basin in California. Field data on water table
elevations were inverted using a hybrid computer to arrive
at spatial variation of aquifer transmissivity. Using a
heuristic approach, estimates were also made of storage
coefficient, and the boundary of the aquifer.
Emsellem and de Marsily (1971) addressed the inverse
problem for steady state flow in an aquifer using the
digital computer. The thrust of their automatic solution
method was that the inverse problem, which is normally
poorly-posed, becomes amenable to solution if available
physical information about the structure of the system is
incorporated into the solution process. At present, inverse
models are mainly used in groundwater hydrology, directed at estimating the spatial distribution of hydraulic
conductivity. The inversion is carried out assuming the
flow system to be under steady-state. To overcome the
difficulty stemming from paucity of data, one may use
any available geological or hydrological knowledge of the
field system to precondition the inversion process. This
methodology, referred to as the Bayesian approach, was
introduced during the 1970s by Delhomme (1979), Neuman and Yakowitz (1979), and others.
A related problem concerns the dependence of the
magnitude of the parameter, hydraulic conductivity, on
scale. This parameter is reasonably understood physically
in terms of impelling and resistive forces on the scale of
laboratory samples of sand, clay, silt, and the like.
However, when one wishes to use the parameter for large
heterogeneous masses of earth materials, its physical
meaning becomes less clear. Yet, the parameter is practically needed on the large scale to mathematically simulate very large groundwater systems. The simple way to
handle this is to use various types of mean values
(arithmetic, harmonic, or geometric means) as appropriate. The phrase upscaling is sometimes used to express
the methods used to describe the generation of parameters
for large-scale models from small-scale information. Such
upscaling of hydraulic conductivity can be useful in the
case of steady-state systems, but may only be of limited
Hydrogeol J (2005) 13:724

values in transient systems. The serious challenge in


simulating the flow behavior of large heterogeneous
systems is to define a set of large-scale average parameters that will somehow preserve important small-scale
effects.
The inverse problem, as well as the problem of upscaling are treated as problems in statistics, or in stochastic processes, as applied to steady-state, heterogeneous systems. Because of the mathematical complexity and
abstract nature of statistics and stochastic processes, the
mathematical symbolism can often be overwhelming and
mask a physical grasp of the results generated by these
models.
Another important problem in numerical modeling is
the uncertainty associated with predictive modeling. This
is because of the errors and ignorance inherent in the
geometrical description, the hydraulic parameters, and the
initial conditions that are used in the calculations. In
traditional engineering practice, such uncertainties are
taken into account through factors of safety incorporated
into the design, or through sensitivity analysis. A more
mathematical approach to handle uncertainty in prediction is to quantify the magnitude uncertainty through
Monte Carlo simulations, by treating the governing of
groundwater flow as a diffusion process with parameters
that are random variables, whose magnitudes are expressed in terms of a mean and a variance.
During the 1960s, when the first numerical models
appeared, there was great anticipation about their ability
to solve a large number of practical problems in hydrogeology. This excitement has since undergone moderation. It is now recognized that numerical models can only
be valuable in providing insights into the potential behavior of complex hydrogeological systems, and to test
alternate hypotheses to better understand observed phenomena. Considering the inaccessibility of the Earths
subsurface, the pervasive heterogeneity on many spatial
scales, the strong interactions among fluid flow, deformation, heat flow, and chemical interactions, and lack of
knowledge of future forcing functions, it will not be
prudent to assume that numerical models will predict the
future with confidence, even with the availability of the
most powerful computing machines.

Near-surface hydrogeological processes


At the turn of the 21st century, processes of the Earths
near-surface and the human habitat are increasingly engaging the attention of earth scientists. Limnologists,
aquatic ecologists, hydrologists, geomorphologists, climatologists, and landscape managers are coming together,
reflecting emerging concerns about human impacts on the
environment. New disciplines such as biogeochemistry,
ecohydrology, and bioclimatology are emerging, all of
which have strong connections to hydrogeology. All share
a common goal of understanding the interconnected hydrological, erosional and nutrient cycles which sustain all
life. The Earths near-surface, where these disciplines
DOI 10.1007/s10040-004-0422-5

19

intersect, includes the soil, the vadose zone, riparian and


hyporheic habitats, wetlands and agricultural lands.
The vadose zone, topped by the soil, mediates between
the atmosphere and the water table. The hydraulic-divide
between the upward-migration of moisture due to evapotranspiration, and gravity-driven downward flow fluctuates in the vadose zone. Chemical weathering, breakdown of organic matter, fixation of nitrogen, and other
chemical reactions so vital for the nutrient cycle largely
occur in the vadose zone, strongly influenced by the
presence of water, air, and microorganisms. Finally, water
that recharges the groundwater reservoir, and the contaminants released at the land surface have first to pass
through the vadose zone. The physical, chemical and biological processes that occur here are of fundamental
geological interest. Here pedology and geology converge.
Lakes, streams, and wetlands, constitute aquatic
ecosystems. The riparian and the hyporheic zone represent, respectively, the transition between the aquatic environment of a stream and the adjoining land, the local
groundwater. Microbial organisms and invertebrates of
the aquatic environment demonstrate adaptation to the
physical and chemical attributes of pore water in these
zones. Here, biogeochemistry, ecology and hydrogeology
come together. Likens and Bormann (1977), in presenting
their findings on a forested ecosystem point out that data
on hydrology is of paramount importance in understanding the biogeochemistry of an ecosystem because water is
a chemical solvent, a catalyst, and a transporting agent.
Euliss et al. (2004) advance the concept of a wetland
continuum, which allows wetland managers, scientists,
and ecologists to simultaneously consider the influence of
climate and hydrologic setting on wetland biological
communities.
Irrigated agriculture around the world has profoundly
impacted the human habitat. One major consequence of
prolonged irrigation is the degradation of soil fertility and
groundwater quality due to water logging. For more than
a century, irrigation engineers have focused attention on
overcoming this problem on a local scale, with the primary objective of maintaining productivity through efficient water use, export of water through tile drains, and
soil amendments. Since export of accumulated salt is not
only expensive and is subject to serious environmental
consequences, the management of irrigated agriculture is
an important issue that needs to be addressed on the scale
of a regional hydrogeological system. On this scale,
pedology, soil geomorphology, agriculture, ecosystems,
and water quality are all rationally integrated.

Groundwater and society


Water occupies a central role in the history of most cultures. Not surprisingly, attitudes towards water management around the world are influenced by local traditions.
Social dimensions of water, therefore, constitute an invaluable adjunct to a scientific understanding of hydrogeological systems. Through a study of the history of the
Hydrogeol J (2005) 13:724

native peoples of the western hemisphere, Back (1981)


provided perspectives on how water resources played a
significant role in the growth as well as the decline of the
early civilizations of North and South America.
Water is often an object of contention between those
who seek its control, and the larger community that depends on it for sustenance. Understandably, litigation for
water has been part of the American society over the past
two centuries. During the nineteenth century, American
courts believed that groundwater was a mysterious phenomenon. For example, the Supreme Court of Ohio
(1861) held that groundwater is so secret, occult, and
concealed that an attempt to administer any set of legal
rules in respect to [it] would be involved in hopeless
uncertainty. This perception of mystery began to change
at the beginning of the 20th century with the systematic
gathering of scientific data from around the country by
the U.S. Geological Survey, and with common citizens
taking note of the effects of groundwater overdraft on the
environment. Over the past few decades, the courts in
America have recognized that groundwater systems can
be scientifically understood in terms of causeeffect relationships, and that such scientific knowledge should
form a rational basis of adjudication.
In the American west it was believed during the late
19th century that the overlying landowner had unlimited
right to all groundwater below ones land and that there
was no limit to the quantity of water that may be extracted
from a well. However, this view became unsustainable
with marked declines in well productivity during periods
of drought. In an influential decision, the Supreme Court
of California (1903) held that burdens due to declined
groundwater yields should be proportionately borne by
overlying land owners. This principle is known as correlative rights.
With the invention of the deep-well turbine pump in
California around 1907, and the availability of electric
power, groundwater became an inexpensive alternative to
surface water for irrigation throughout the American
west. Encouraged by the potential for economic prosperity, and poorly defined ownership rights, many western states witnessed over-production of groundwater.
Within decades, this led to large declines in water levels,
disappearance of artesian flowing wells, land subsidence,
earth fissuring, and saltwater intrusion in coastal aquifers.
Clearly, groundwater withdrawal had to be controlled.
But, the early laws were primarily intended to foster
economic growth, and did not show awareness of the
finiteness of the resource. Planned groundwater development required enactment of new laws.
In the United States, water management is a state
subject, except in regard to water rights of Native
Americans. The states own all water within their boundaries in trust for the people. In regard to Native Americans, the Federal government acts as the guardian of their
water rights. That the states own all water in trust for the
people is based on historic Roman law traditions, and is
referred to as the doctrine of public trust (Narasimhan
2003). This doctrine is now being used as a basis by many
DOI 10.1007/s10040-004-0422-5

20

states to mandate integrated management of surface water


and groundwater on a basin-wide scale. Individuals, corporations, and municipalities are required to obtain permits for specific beneficial use of water, with the understanding that water will not be wasted. The permit is
usufructuary in the sense that the resource itself will not
be unduly damaged in the process of the beneficial use.
State and federal environmental protection laws have
been set in place to protect groundwater bodies from
chemical degradation due to human activities. Although,
during the 1980s, some courts extended the public trust
doctrine to ecological values, its application to groundwater is still being contested in courts by those who acquired riparian and appropriation rights during the 19th
and early 20th century.
In the wake of the strong environmental movement of
the 1960s, there exist two opposing perceptions about
groundwater. Those with acquired water rights would like
having groundwater treated as private property. This is
especially true in California, where some farmers own
very large tracts of land and associated water rights. On
the other hand, many hydrogeologists, ecologists, and
environmentalists would like to see groundwater and
surface water treated as a single resource, and subjected to
integrated management on a basinal scale. On a scientific
level, it is becoming clear that the groundwater reservoir
must be so managed as to play the role of a buffer that can
be relied upon to tide over water deficiency during periods of drought. Many western states, such as Arizona,
Nebraska, New Mexico, and Texas have already moved in
the direction of asserting the States responsibility to
manage surface water and groundwater as a single integrated resource.
An interesting recent development is that institutions
and agencies involved with water have made public education a part of their activities. As a result, the common
citizen is becoming increasingly aware of the nature of
the hydrological cycle, and within it, the groundwater
system. This public literacy, it appears, will eventually
lead to a wise management of surface water and
groundwater, duly considering the needs of humans and
of ecosystems.

Future: a speculation
There are many facets to hydrogeology: the science itself,
its applied benefits, the forces that drive its growth, and
its relation to society. Speculation about its future,
therefore, must be based on expectations about how these
may change with time. Hydrogeology constitutes the
study of geological processes influenced by water. The
motivation to pursue it may be mere intellectual curiosity,
or may stem from a desire to solve problems of interest to
society, or both. In our contemporary society, hydrogeology research is driven by governmental funding, and to
a lesser extent, industrial support. Society is increasingly
becoming aware of groundwater, as freshwater is rendered scarce by the needs of a technological society.
Hydrogeol J (2005) 13:724

The observational base of hydrogeology is expanding


at a phenomenal pace. We now observe the Earths history, in real time, at very high spatial and temporal resolutions. There is fervent expectation that these data can
help predict and control the behavior of hydrogeological
systems for much benefit to society. How realistic is this
expectation?
The science of hydrogeology is based on principles of
the physical sciences. Physical sciences strive for precise
measurement, detailed description, and reliable prediction
of the behavior of the physical world. Inspired by this,
modern technology pursues the goal of controlling the
physical world for societal benefit. Nonetheless, the inaccessibility of the Earths subsurface, pervasive heterogeneity on many spatial and temporal scales, and the
limitations inherent in mathematical tools that seek to
quantify hydrogeological processes cast serious doubt on
the ability of the physical sciences to precisely predict and
control the behavior of these systems. Moreover, the behavior of living things, stemming from their ability to
adapt to their changing environment complicates human
aspirations of prediction and control of hydrogeological
systems.
Hydrogeology is a historical science. It is concerned
with comprehending why the earth is what it is today
because of what happened in the past. In this role, hydrogeology is enjoying enormous success. Experience has
shown that, despite this success, predicting the long-term
future of hydrogeological systems is very difficult, or
even impossible, because of limitations inherent in our
models, and the unpredictability of the forces that drive
the earth.
The ability to collect data of unprecedented resolution
on the one hand, and the severe limitations that confront
prediction and control on the other, raise a serious issue.
The future of hydrogeology will clearly depend on the
expectations with which financial support will be forthcoming for research. Although a small proportion of total
funding may be designated for uncommitted scientific
inquiry, a major portion will be allotted with expectations
of applied benefit. Some important questions arise.
Should society continue emphasis on gathering more and
more data, or should resources be diverted to better
management of hydrogeological systems based on what
has already been established? What should be the motivation for collecting data from hydrogeological systems?
Should it be the timely detection of unacceptable consequences of human actions, or should it aim at more efficient control of such systems to serve human needs?
Funding agencies face a dilemma: Should it be assumed
that evolving technology will continue to overcome Natures constraints and enhance human ability to predict
and control hydrogeological systems, or should it be assumed that these systems cannot be precisely predicted or
controlled. Each of these assumptions will demand a
different type of research output. The former course
would take a path of increased exploitation of hydrogeological systems, while the latter will lead to adaptive
management, constraining the use of such systems within
DOI 10.1007/s10040-004-0422-5

21

definite bounds. The former would anticipate new discoveries and inventions. The latter, careful monitoring of
existing systems for timely detection of potentially unacceptable impacts.
Over 50 years ago, Hubbert studied the growth and
decline of a number of non-renewable resources such as
minerals, coal, and hydrocarbons, and vigorously argued
that there are not enough natural resources in the Earth to
sustain significant economic growth indefinitely into the
future. His view was that there already exists enough
scientific know-how to pursue a more realistic long-term
goal of steady or very slight economic growth over long
periods of time (Hubbert 1973). It is of value to extend
Hubberts analysis to hydrogeological systems, and understand how they degrade with time. This task is more
complicated because many hydrogeological systems are
partly renewable, and the notion of degradation has to
address not only water itself, but also the soil and other
components of the ecosystem that sustain flora and fauna.
An emerging trend in the earth sciences is the recognition that the hydrological, the erosional the geochemical
and the nutritional cycle are intimately interlinked. These
cycles sustain life, and are, in turn, influenced by it. It is
in this context that the future of hydrogeology may be
rationally speculated upon. Just as hydrogeology found its
links to geochemistry, tectonics, and petroleum geology
during the 1950s, so also, hydrogeology is currently discovering new connections to biogeochemistry, ecohydrology, bioclimatology, and other emerging areas of
study. In so far as water is the principal agent that
transports energy and matter in the lithosphere, hydrogeology has the potential to bring all these disciplines
together through the framework of regional groundwater
flow systems. As attention shifts to understanding the lifesustaining cycles, it is likely that distinctions among fields
such as hydrogeology, biogeochemistry, ecohydrology,
may become less important.
Substantial knowledge now exists about the physical
and chemical processes that govern hydrogeological systems, including fluid flow, deformation, energy transfer,
and chemical reactions. One would like to believe that the
conceptual mathematical knowledge that exists in these
areas probably exceeds our current abilities to make field
measurements. At the same time, experience with the
modeling of hydrogeological systems have shown that the
equations of mass and energy conservation, transport, and
thermodynamics on which the models are based, are
themselves only approximations to the observable system.
As a result, considerable difficulties exist in making the
observed system fit even the best available mathematical
models, and adequately characterize the modeled system.
The maximum potential for new knowledge is likely to be
in two areas; the role of microbes in hydrogeological
systems, and marine hydrogeology.
Over the past few decades, substantial evidence has
accumulated showing that microbial organisms play a
profound role in many hydrogeological processes from
chemical weathering, soil formation and petroleum genesis to nutrient cycling. Yet, the physical and chemical
Hydrogeol J (2005) 13:724

mechanisms associated with their actions is largely unknown. Additionally, microbes possess life, and they
respond, adapt, and exhibit a will to survive within
their changing environment. These attributes are not
amenable to description by physical laws that constitute
the basis for describing the behavior of inanimate things.
Marine hydrogeology is a young field, with impressive
international cooperation, and driven largely by questions
aimed at understanding how the Earth functions. In this
field, research questions are posed in a manner that the
answers must account for physical, chemical, and biological observations on a variety of space and time scales.
Compared with continental hydrogeology, the marine
counterpart is more resource intensive, and larger in scale.
As a consequence, the future of marine hydrogeology will
very much depend on the ability of funding agencies to
invest in what is heavily basic science, with less emphasis
on immediate applied benefit.
In applied hydrogeology, a topic of great interest is
sustainable management of groundwater resources. Here,
sustainable management implies the maintenance of stable water supplies for society, and assuring the longevity
of the resources for future generations. Longevity, in this
context, may significantly exceed traditional time scales
of engineering decision-making, and extend to thousands
of years. There are many components to this task: coordinated use of surface water and groundwater in such a
way that the latter stores excess supplies during wet years,
and acts as a buffer during droughts; maintenance of riparian, hyporheic, and other habitats, minimization of
chemical contamination of fresh groundwater bodies, and
restoration of aquifers and habitats that have been impaired by prior human actions. Clearly, managing water
involves more than just water. Vigorous industrial production is inevitably accompanied by stressing the productivity of hydrogeological systems to their limits, and
contaminating them chemically. For a long time, the
impacts of these degradations were simply ignored, in
favor of the purported economic benefits. More recently,
efforts have been made to look at resource degradation in
terms of monetary benefitcost analysis. However, there
are concerns that these analyses are far from adequate in
accounting for the value of the lost resource to future
generations. Judicious management of groundwater systems will, in the future, entail far more than mere science.
Many challenges lie ahead in combining human values
with applied hydrogeology.
Speculation about the future of hydrogeology cannot
ignore the role of computers and mathematical models.
During the 1960s, when computer models for analyzing
groundwater systems made their appearance, there was
great excitement about their potential ability to predict the
future behavior of groundwater systems, and thus to
manage them for great benefit and profit. Despite explosive developments in computing power since then, current
expectations about the ability of the computers to predict
the future are quite subdued. It is now recognized that
mathematical models are valuable tools for looking at
different scenarios, and testing alternative hypotheses,
DOI 10.1007/s10040-004-0422-5

22

rather than being predictive tools. Modern computing


machines provide phenomenal possibilities for storing
and retrieving large amounts of data, and portraying them
in useful ways.
One inherent limitation of mathematical models of
hydrogeological systems is that boundary conditions that
force the systems to change are external to the models,
and have to be made available independently. Common
practice is to prescribe these boundary conditions based
on limited past experience. However, emerging knowledge about paleohydrology has added a new dimension to
the difficulty of modeling hydrogeological systems. Recent developments based on the study of ice cores from
Antarctica and elsewhere (Alley et al. 2003) suggest that
drastic climatic changes from glacial to inter-glacial climates can occur over a period of just a few decades, and
that droughts on a continental scale can occasionally
persist for a century or more. The forces that cause such
changes are as yet unknown. Suppose large tracts of the
United States or Europe are subjected to droughts spanning many decades. What would be the impact on the
hydrogeological systems that sustain these regions? How
will the technological societies adjust themselves? It is
impossible to answer these questions because human behavior under conditions of extreme stress cannot be predicted. What can be stated, though, is that the worldwide
response of societies under such contingencies could either be peaceful adjustment, or violent destruction. Water
could conceivably play a global role that will rival the
role of energy. Should society choose to adapt itself to the
constraints of a finite nature and the unpredictability of
forcing functions, then monitoring of sustained hydrogeological systems will become an integral part of managing such systems. The purpose of monitoring would be to
foresee, in a timely way any potential degradation of the
system in unacceptable ways so that adequate preventive
actions can be initiated. Here monitoring is to be interpreted broadly to include physical, chemical, and biological attributes.
Two centuries after the industrial revolution, humans
are beginning to comprehend the importance of the hydrological, nutritional, geochemical, and erosional cycles
that are vital for the sustenance of all life. The future of
hydrogeology depends on how humans choose to live
within the constraints of these cycles. Philosophically,
hydrogeology has the potential to provide a unifying
framework to bring together a number of disciplines
within the earth sciences, biological sciences, and social
sciences. Much will depend on how hydrogeologists
perceive their own identity, and how the various funding
agencies perceive the existence of a technological society
within a finite and bounded Earth.
Acknowledgments I am grateful to Stanley N. Davis for many
informative exchanges over the years on the history of hydrogeology. I thank Ghislain de Marsily, David E. Prudic, and Clifford I.
Voss for a critical reading of the manuscript and many constructive
criticisms. This work was partly supported by funds from the
Agricultural Extension Service, through the Division of Natural
Resources, University of California
Hydrogeol J (2005) 13:724

References
Alley RB, Marotzke J, Nordhaus WD, Overpeck JT, Peteet DM,
Pielke Jr RA, Pierrehumbert RT, Rhines PB, Stocker TF, Talley
LD, Wallace JM (2003) Abrupt Climate Change, Sci 299:2005
2010
Back W (1960) Origin of hydrochemical facies of groundwater in
the Atlantic Coastal Plain. Rept. 21st Session, Int. Sci. Cong.,
Copenhagen, Part 1, pp 8795
Back W (1981) Hydromythology and ethnohydrology in the New
World. Water Resour Res 17:257282
Back W, Freeze RA (1983) Chemical hydrogeology. Hutchinson
Ross Publishing Co., Stroudsberg, PA, 416 pp
Barenblatt GI, Zheltov IP, Kochina IN (1960) Basin concepts in the
throry of seepage of homogeneous liquids in fissured rocks.
J Appl Math 24:12861303
Bear J (1960) The transition zone between fresh and salt waters in
coastal aquifers. Ph.D. Dissertation, Civil Engineering, University of California at Berkeley, 139 pp
Berry F, Hanshaw BB (1960) Geologic field evidencd suggesting
membrane properties of shales. 21st Int. Geol. Congress,
Copenhagen, Abstracts, p 209
Biot MA (1941) General theory of three-dimensional consolidation.
J Appl Phys 12:155164
Bishop AW, Blight GE (1963) Some aspects of effective stress in
saturated and partly saturated soils. Geotechnique 13:177197
Boulton NS (1954) Unsteady radial flow to a pumped well allowing
for delayed yield from storage. Int Assoc Sci Hydr Binder
4:472477
Boussinesq MJ (1868) Mmoire sur linfluence des frottements
dans les mouvements rgulaires des fluides. (Memoir on the
influence of resistance in laminar flow in fluids.) J de Mathematique pures et aplliques, Gautheir-Villars, Paris 13:377424
Bredehoeft JD (1967) Response of well-aquifer systems to earth
tides. J Geophys Res 72:3075308
Bredehoeft JD, Back W, Hanshaw BB (1982) Regional groundwater flow concepts in the United States: historical perspective.
In: Narasimhan TN (ed) Recent trends in hydrogeology. Special
Paper 189, Geological Society of America, pp 297316
Buckingham E (1907) Studies on the movement of soil moisture.
Bureau of Soils, U.S. Department of Agriculture, Bull. No. 38,
61 pp
Bullard EC, Maxwell AE, Revelle R (1956) Heat flow through the
deep sea floor. In: Landsberg HE (ed) Advances in geophysics,
vol 3. Academic Press Inc., New York, pp 153181
Carlston CW (1963) An early American statement of the Badon
Ghyben-Herzburg principle of static fresh-water/salt-water
balance. Amer J Sci 261:8891
Cathles LM, Smith AT (1983) Thermal constraints on the formation of Mississippi Valley-type leadzinc deposits and their
implications for episodic basin dewatering and deposit genesis.
Econ Geol 78:9831002
Chebotarev II (1955) Metamorphism of natural waters in the crust
of weathering. Geochim Cosmochim Acta 8:2248,137
170,198212
Clough RW (1960) The finite element method in plane stress
analysis. In: Second Conference on Eletronic Computation,
Amer. Soc. Civil Eng., The Pittsburg Section, pp 345377
Cooper HH Jr, Bredehoeft JD, Papadopulos IS, Bennett RR (1965)
The response of a well-aquifer system to seismic waves.
J Geophys Res 70:39153926
Darton NH (1896) Preliminary report on artesian waters of a portion of the Dakotas. In: 17th Annual Rept., U.S. Geol. Surv.,
pp 609694
Darcy H (1856) Dtermination des lois dcoulement de leau 
travers le sable. In: Les Fontaines Publiques de la Ville de
Dijon, Victor Dalmont, Publisher, pp 590594
Davis SN (1989) What is hydrogeology. Ground Water 27:143144
Davis SN (1999) Humboldt, Arago, and the temperature of
groundwater. Hydrogeol J 7:501503
Davis SN, Davis AG (1997) Saratoga Springs and early geochemistry in the United States. Ground Water 35:347356
DOI 10.1007/s10040-004-0422-5

23
Delhomme JP (1979) Spatial variability and uncertainty in
groundwater flow parameters, a geostatistical approach. Water
Resour Res 15:269280
Deming D (2002) Origin of oceans and continents. Int Geol Rev
44:137152
Du Commun J (1828) On the cause of freshwater springs, fountains
& c. Amer J Sci 1st Ser 14:174176
Dupuit AJEJ (1857) Mmoir sur le mouvement de leau  travers
les terrains permables. Comptes Rendus Hebdomadaires des
Sances de lAcadmie des Sciences (Paris) 45:9296
Dupuit AJEJ (1863)
tudes thoritiques et pratiques sur le mouvement des eaux dans les canaux dcouverts et a travers les
terrains permables, Deuxime
dition. Dunod, Paris, 304 pp
Eakin TE (1966) A regional interbasin ground-water system in the
White River area, southeastern Nevada. Water Resour Res
2:251271
Emsellem Y, de Marsily G (1971) An automatic solution for the
inverse problem. Water Resour Res 7:12641283
Euliss NH, La Baugh JW, Frederickson LH, Mushet DM, Laubhan
MK, Swanson GA, Winter TC, Rosenberry DO, Nelson RD
(2004) The wetland continuum: a conceptual framework for
interpreting biological studies. Wetlands 24:448458
Evans DM (1966) The Denver area earthquakes and the Rocky
Mountain Arsenal Disposal Well. Mountain Geologist 3:2336
Fourier JBJ (1822) Thorie Analytique de la Chaleur, F. Didot,
Paris
Freeman VM (1968) People-land-water: Santa Clara Valley and
Oxnard Plain, Ventura County. L.L. Morrison, Los Angeles,
227 pp
Freeze RA (1966) Theoretical analysis of regional groundwater
flow. Ph.D. Dissertation, University of California at Berkeley
Freeze RA (1985) Historical correspondence between C.V. Theis
and C.I. Lubin. EOS, Trans. Amer. Geophys. Union 66:4142
Gardner W, Israelsen W, Edlefsen NE, Clyde H (1922) The capillary potential function and its relation to irrigation practice,
Abstract. Phys. Rev. Ser. II 20:199
Garrels RM (1960) Mineral equilibria at low temperature and
pressure. Harper and Brothers, New York, 254 pp
Gelhar LW, Gutjahr AL, Naff RL (1979) Stochastic analysis of
macrodispersion in a stratified aquifer. Water Resour Res
15:13871397
Ghyben WB (1888) Nota in verband met de voorgenomen putboring nabij Amsterdam, Tijdschgrift van Let Koninklijk Inst.
van Ing. (Note in connection with the planned borehole drilling
near Amsterdam, Jour. Royal Inst. Engrs.)
Gilbert GK (1890) Lake Bonneville. Mem US Geol Surv 1:1438
Hall HP (1954) A historical review of investigations of seepage
toward wells. J Boston Soc Civil Eng 41:251311
Helgeson HC (1968) Evaluation of irreversible reactions in geochemical processes involving minerals and aqueous solutions
I. Thermodynamic relations. Geochem Cosmochim Acta
32:853877
Hem JD (1959) Study and interpretation of the chemical characteristics of natural water. U.S. Geol. Surv. Water Supply Paper
1473, 269 pp
Hem JD, Cropper WH (1959) Survey of ferrous-ferric chemical
equilibria and redox potentials. U.S. Geol. Surv. Water Supply
Paper 1459-A
Herzberg A (1901) Die Wasserversorgung einiger Nordseebarder.
Jour Gasbeleucht Wasserversorg 44:815819 (The water
treatment of Northseabarder, Jour. Gas Illumination and Water
Treatment.)
Hubbert MK (1940) The theory of ground-water motion. J Geol
48:785944
Hubbert MK (1953) Entrapment of petroleum under hydrodynamics conditions. Bull Amer Assoc Petroleum Geol 37:19542026
Hubbert MK (1973) Survey of world energy resources. Can Mining
Meta Bull 66:3753
Hubbert MK, Willis DG (1957) Mechanics of hydraulic fracturing.
Amer Inst Min Eng Petroleum Trans 210:153168
Hubbert MK, Rubey W (1959) Role of fluid pressure in mechancis
of overthrust faulting: Pt. 1. Mechanics of fluid-filled porous
Hydrogeol J (2005) 13:724

solids and its application to overthrust faulting. Bull Geol Soc


Amer 70:115166
von Humboldt A (1844) Cosmos: a sketch of a physical description
of the Universe, vol 1. Harper and Brothers Publishers, New
York (translated from German by E.C. Ott, published in 1897)
Jacob CE (1940) Flow of water in an elastic artesian aquifer. Trans
Amer Geophys Union 21:574586
Jacob CE (1946) Radial flow in a leaky artesian aquifer. Trans
Amer Geophys Union 27:198208
King FH (1892) Observations and experiments on the fluctuations
in the level and rate of movement of ground water on the
Wisconsin agricultural experiment station farm, and Whitewater, Wisconsin. US Weather Bur Bull 5:6769
Klute A (1952) A numerical method for solving the flow equation
for unsaturated soil. Soil Sci 20:317320
Lee CH (1912) An intensive study of the water resources of a part
of the Owens Valley, California. US Geol Surv Water Supply
Paper 294, 135 pp
Likens GE, Bormann H (1977) Biogeochemistry of a forested
ecosystem. Springer, Berlin Heildelberg New York, 159 pp
Lucas J (1877) The Chalk River system. Proc Inst Civil Eng 47:70
167
MacNeal RH (1953) An asymmetric finite difference network.
Quart Appl Math 2:295310
Marriott E (1686) Traite du mouvement de eaux et des autres
corps fluides
Mather J (2001) Joseph Lucas and the term hydrogeology.
Hydrogeol J 9:413415
Meinzer OE (1928) Compressibility and elasticity of artesian
aquifers. Eco Geol 23:263291
Meinzer OE (1934) Hydrologythe history and development of
ground-water hydrology. Wash Acad Sci 24:632
Meinzer OE (1937) Land subsidence caused by pumping. Eng
News Rec 18:715
Meyboom P (1962) Patterns of groundwater flow in the prairie
profile. In: Proc. Hydr. Symp. No. 3 Groundwater, Water Resources Branch Department of Northern Affairs and National
Resources and Geological Survey of Canada, pp 320
Narasimhan TN (1975) A unified numerical model for saturatedunsaturated groundwater flow. Ph.D. Dissertation, University of
California at Berkeley, 244 pp
Narasimhan TN (1986) Evolution of the notion of time in hydrogeology. EOS, Trans. Amer. Geophys. Union 67:789790
Narasimhan TN (2003) A finite world, earth sciences, and public
trust. Ground Water 41:1114
Neuman SP, Yakowitz S (1979) A statistical approach to the inverse problem of aquifer hydrology. Part I.Theory. Water Resour Res 15:845860
Palmer C (1911) The geochemical interpretation of water analyses.
U.S. Geol Surv Bull 479:31
Perrault P (1674) De Lorigine des fontaines, Paris. Translation: on
the origin of springs, Hafner, New York, by Aurele La Rocque,
1967
Philip JR (1956) The theory of infiltration, 1. The infiltration
equation and its solution. Soil Sci. 83:345357
Piper AM (1944) A graphic procedure in the geochemical interpretation of water-analyses. Trans Amer Geophys Union
25:914923
Pratt WE, Johnson DW (1926) Local subsidence of the Goose
Creek oil field. J Geol 34(part 1):577590
Pruess K, Narasimhan TN (1982) On fluid reserves and the production of superheated steam from fractured vapor-dominated
geothermal reservoirs. J Geophys Res 87:93299339
Raleigh CB, Healy JH, Bredehoeft JD (1976) An experiment in
earthquake control at Rangely. Colorado Sci 191:12301237
Rappleye HS (1933) Recent areal subsidence found in releveling.
Engin News Rec 110:845
Renick BC (1924) Base exchange in ground water by silicatesas
illustrated in Montana. U.S. Geol. Surv. Water Supply Paper
520-D
Richards LA. (1931) Capillary conduction of liquids through porous mediums. Physics 1:318333
DOI 10.1007/s10040-004-0422-5

24
Rogers GS (1917) Chemical relationships of the oil-field waters in
San Joaquin Valley, California. U.S. Geol. Surv. Bull. 653, 119
pp
Russell IC (1885) Geological history of Lake Lahontan, a quarternary lake of northwestern Nevada. Mem US Geol Surv 11:1
288
Scheidegger AE (1954) Statistical hydrodynamics in porous media.
J Appl Phys 25:9941001
Seed HB, Lee KL, Idriss IM, Makdisi FI (1973) Analysis of the
slides in the San Frnando dams during the earthquake of Feb. 9,
1971. Report No. EERC 73-2, College of Engineering, University of California at Berkeley, 150 pp
Simpson DW, Narasimhan TN (1990) Inhomogeneities in rock
properties and their influence on reservoir induced seismicity.
Gerlinds Beitrage zur Geophysik 99:205219
Skempton AW (1954) The pore pressure coefficients A and B.
Geotechnique 4:143147
Slichter CS (1899) Theoretical investigation of the motion of ground
waters. Ann. Rept. 19, Part II, U.S. Geol. Surv., pp 295384
Snow DT (1965) A parallel plate model for fractured permeable
media. Ph.D. Dissertation, Engineering Science, University of
California, Berkeley, 331 pp
Snow DT (1969) Anisotropic permeability of fractured media.
Water Resour Res 5:12731289
Supreme Court of California (1903) Katz vs. Walkinshaw, 141 Cal.
116, 74 P. 766
Supreme Court of Ohio (1861) Frazier vs. Brown, 12 Ohio St. 294
Terzaghi K (1923) Die Berechnung der Durchl ssigkeitsziffer des
Tones aus dem Verlauf der hydrodynamischen Spannungserscheinungen [The calculation of permeability number of the
clay out of the process of the hydrodynamic phenomenon
tension]. Sitz. Akad. Wissen. Wien. Math-naturw. Kl., Part Iia,
32:125138

Hydrogeol J (2005) 13:724

Theis CV (1935) The relation between the lowering of the lowering


of the piezometric surface and the rate and discharge of a well
using ground water storage. Trans Amer GeophysUnion
16:519524
Toth J (1962) A theoretical analysis of groundwater flow in small
drainage basins.In: Proc. Hydr. Symp. No. 3 Groundwater,
Water Resources Branch Department of Northern Affairs and
National Resources and Geological Survey of Canada, pp 75
96
Toth J, Millar RF (1983) Possible effects of erosional changes of
the topographic relief on pore pressures at depth. Water Resour
Res 19:15851597
Tyson HN, Weber EM (1964) Groundwater management for the
nations futurecomputer simulations of groundwater basins.
Amer Soc Civil Eng, J Hydr Div 90(HY4):5977
Vemuri V, Karplus WJ (1969) Identification of nonlinear parameters of groundwater basins by hybrid computation. Water
Resour Res 5:172185
Warren JE, Root PJ (1963) The behavior of naturally fractured
reservoirs. Soc Pet Eng J 245254
White DE (1965) Saline waters of sedimentary rocks. Amer Assoc
Petroleum Geol Mem 4:342366
White WB (1988) Geomorphology and hydrology of Karst terrains.
Oxford University Press, New York, 464 pp
Wilson EL (1968) The determination of temperatures within mass
concrete structures. In: Rept. 6817, Structural Engineering
Laboratory, Department of Civil Engineering, University of
California at Berkeley, 33 pp
Winograd IJ (1962) Interbasin movement of ground water at the
Nevada Test Site, Nevada. In: Short Papers in Geology and
Hydrology: U.S. Geol. Surv. Prof. Paper 450-C, C108C111

DOI 10.1007/s10040-004-0422-5

You might also like