Fraccionamiento Isotopico Gard
Fraccionamiento Isotopico Gard
Fraccionamiento Isotopico Gard
J. R. Gat
Department of Environmental Sciences and Energy Research, Weizmann
Institute of Science, 76100 Rehovot, Israel
ABSTRACT
Changes of the isotopic composition of water within the water cycle provide
a recognizable signature, relating such water to the different phases of the cycle.
The isotope fractionations that accompany the evaporation from the ocean and
other surface waters and the reverse process of rain formation account for the most
notable changes. As a result, meteoric waters are depleted in the heavy isotopic
species of H and O relative to ocean waters, whereas waters in evaporative systems
such as lakes, plants, and soilwaters are relatively enriched. During the passage
through the aquifers, the isotope composition of water is essentially a conservative
property at ambient temperatures, but at elevated temperatures, interaction with
the rock matrix may perturb the isotope composition. These changes of the isotope
composition in atmospheric waters, surface water, soil, and groundwaters, as well
as in the biosphere, are applied in the characterization of hydrological system as
well as indicators of paleo-climatological conditions in proxy materials in climatic
archives, such as ice, lake sediments, or organic materials.
1. INTRODUCTION
The basic tenet of the hydrologic cycle as expounded by the preacher, that
“into the sea all the rivers go and yet the sea is never filled, and still to their
goal the rivers go” (The Jerusalem Bible, English translation, 1970), remains
valid until today. However, the simple concept of a steady-state one-cycle
225
0084-6597/96/0515-0225$08.00
March 16, 1996 8:49 Annual Reviews GAT-8 AR10-08
226 GAT
Figure 1 The hydrologic cycle, in relative flux units (100 units equals the marine evaporation
flux). Figures have been rounded up. (Adapted from Chow 1964.)
as in ice cores, lake and ocean sediments, and of plant material, have proven
useful for paleoclimate reconstructions.
The abundance of the isotopes 2 H (deuterium, also marked as D) and 18 O in
ocean waters is close to D/H = 155.95 × 10−6 (Dewit et al 1980) and 18 O/O
= 2005.2 × 10−6 (Baertschi 1976), respectively, where H and O without a mass
assignment stands for the natural assembly of all isotopic species. Another
natural oxygen isotope, 17 O, occurs at an abundance of about 1/10 that of 18 O,
but because its variations in the water cycle parallels that of the heavier isotope
(at 1/2 the extent) it has not been measured as a rule.
Abundance variations of ± 30% for deuterium and ± 5% for 18 O have been
measured in natural waters (Boato 1961). These variations are only to a neg-
ligible degree the result of nuclear interactions that produce or consume these
isotopes; they rather result primarily from the fractionation of the isotopic wa-
ter species in isotopic exchange reactions (equilibrium isotope effects) or in
chemical or biochemical reactions involving the isotopic species (kinetic and
vital effects), or as a result of different diffusion rates of the isotopic water
molecules (transport effects). In the water cycle, the most significant process
in this respect is that of phase changes, from vapor to liquid or ice and vice
versa.
The ocean, being the largest water reservoir and relatively homogeneous,
was chosen as the reference standard for the δ scale of both oxygen and hydro-
gen isotopes in water samples (Craig 1961a). The δ-(permil) value is defined
as δ‰ = (Rx /Rstd − 1) × 103 , where R is the atom ratio D/H and 18 O/16 O,
respectively. Positive δ values thus signify an enrichment of the heavy iso-
topic species relative to the standard (SMOW); negative values indicate their
depletion. Because the variability in δ(D) is typically larger by almost an order
of magnitude than that of δ(18 O), the usual way of presenting isotope data of
water samples is on a δ(D) vs δ(18 O) diagram with the δ(D) scale compressed
by a factor of ten.
Ever since the earliest measurement of isotopic abundance in the 1930s in
Europe and Japan (Rankama 1954) and by the Chicago school of HC Urey, us-
ing the Nier-McKinney double-inlet, double-collector mass-spectrometer (Nier
1947, Mckinney et al 1950), it has been observed that freshwater samples are
depleted in the heavy isotopes, whereas the residue of evaporating surface wa-
ters such as African lakes are enriched (Craig 1961b). A correlation between the
depletion of the deuterium and oxygen-18 in freshwaters was established; the
locus line of such samples in δ-space was defined as δ(2 H) = 8δ(18 O)+10(‰).
This line is now known as the Global Meteoric Water Line (GMWL).
March 16, 1996 8:49 Annual Reviews GAT-8 AR10-08
228 GAT
also apply to the case where material is removed by diffusion or outflow, e.g. by
effusion through an aperture. In the latter cases the term transport fractionation
factor may be preferred. Unlike true kinetic fractionation factors, which like
the equilibrium factors are strongly temperature dependent, the transport factors
show only slight (positive) temperature coefficients.
The values of the equilibrium fractionation factors (α + ) for the water-vapor
phase transition are given by Majoube (1971) as 1.0098 and 1.084 at 20◦ C and
1.0117 and 1.111 at 0◦ C, respectively, for 18 O and D. For the ice-liquid transition
the values of α + (at 0◦ C) are 1.0035 for 18 O and 1.0208 for D (Arnason 1969).
2.2 The Rayleigh Equations
If material is removed from a (mixed) system containing Ni and N j molecules,
respectively, of two isotopic species and if the fractionation accompanying the
removal process at any instance is described by the unit fractionation factor
α, then the evolution of the isotopic composition in the remnant material is
described by the following equations, assuming N Ni , where Ni + N j = N ,
so that it follows that N ' N j (which holds for the natural isotope abundance
of the lighter elements, e.g. H, N, C, and O):
µ ¶
dR 1 d Ni Ni R
= − = (α − 1) (1a)
dN N dN N N
and
d(ln R)
= (α − 1). (1b)
d(ln N )
Equation 1a can be immediately integrated from an initial condition (R0 , N0 ) to
any given stage, provided α does not change during the course of the process:
µ ¶(α−1)
N
R = R0 = R0 · f (α−1) , (3)
N0
where f = N /N0 is the fraction of the material remaining.
When the isotopic species removed at every instant are in thermodynamics
equilibrium with those remaining in the system, we have the circumstances of
the so-called Rayleigh distillation. Under these conditions, the unit separation
factor will be a thermodynamic equilibrium constant α ∗ , i.e. the vapor pressure
ratio of the isotopic water molecules for the liquid to vapor transitions or the
equilibrium constant of an isotopic exchange reaction.
The isotopic composition R of the remaining material, as a function of f ,
when subject to Rayleigh fractionation, is shown in Figure 2. Obviously, the
isotope composition of the material that is removed at every instance describes
March 16, 1996 8:49 Annual Reviews GAT-8 AR10-08
230 GAT
Figure 2 Isotopic change under Rayleigh, closed, and steady-state conditions for fractionation
processes with a unit fractionation factor of α = 1.01 for initial liquid composition of δ = 0. A, B,
C: Rayleigh conditions for residual water, continuously removed vapor, and total vapor removed,
respectively; D, E: liquid and vapor for two coexisting phase systems; F: liquid approaching steady
state (time axis not in proportion).
March 16, 1996 8:49 Annual Reviews GAT-8 AR10-08
a curve parallel to that of the remaining fraction. The integrated curve, giving
the isotopic composition of the accumulated material thus removed, is also
shown. Material balance considerations require that the isotope content of the
total accumulated amount of the removed material approaches R0 as f → 0.
and also
R2
= α.
R1
Defining f = N1
N1,0
, then
R0
Rf = (3a)
α − f (α − 1)
or in δ(‰) nomenclature
δ0 − ε(1 − f )
δf = . (3b)
α − f (ε/103 )
Unlike in the open system, δ f →0 −→ (δ0 − ε)/α, limiting the degree of isotope
enrichment.
A third model is that of a throughflow system in which the outgoing (frac-
tionated) flux is replaced by an inflow with isotope composition δin . At steady
state, the isotope composition of in- and outflowing fluxes are obviously equal.
If the outflowing flux is related to the reservoir’s isotope composition by a frac-
tionation factor α, then the steady-state isotopic composition in the reservoir
March 16, 1996 8:49 Annual Reviews GAT-8 AR10-08
232 GAT
will be δss = (δin − ε)/α, while the approach to steady-state composition can
be expressed as (Zimmermann 1979)
δss − δ
= exp (−α t/τ ), (4)
δss − δ0
where τ is the characteristic turnover time of the system, given by the volume
to flux ratio (τ = V /F).
Note that for the same unit fractionation factor the resultant evolution of the
isotopic composition is quite different under these different circumstances, as
shown in Figure 2.
In all examples given so far, a constant unit fractionation factor is assumed
to apply throughout the process. This is not, however, always the case. For
example, as will be discussed, the rainout from an air mass is usually the result
of a continuous cooling of the parcel of air concerned; the cooling results in a
change of the unit fractionation factor for the vapor-to-water (or vapor-to-ice)
transition during the evolution of the precipitating system. Although the differ-
ential form of the Rayleigh equation (Equation 1a) still applies, the integration
obviously has to be carried out for a prescribed change of α throughout the
process [as was first demonstrated by Dansgaard (1964)].
Another conspicuous example of a changing “effective” fractionation factor
is that of the evaporation of water from a surface water body into the atmosphere.
In this case, the change is the result of the changing boundary conditions (of
the isotopic gradient, primarily) rather than a change of the unit fractionation
factors themselves.
2.4 Isotope Fractionation Accompanying Evaporation
of Surface Waters
Water-air interaction balances two opposing water fluxes, one upward from the
surface and the other a downward one of atmospheric moisture. At saturation,
i.e. when atmospheric humidity is 100% in terms of the saturation vapor pres-
sure at the liquid surface, this interaction will bring the liquid water and air
humidity into isotopic equilibrium with one another. Such a situation appar-
ently occurs at cloud base between the falling rain droplets and the ascending
air.
When the air is undersaturated, a net evaporation flux results, in which the
rate-determining step is the diffusion of water vapor across the air boundary
layer in response to the humidity gradient between the surface and the fully
turbulent ambient air. Three factors are involved in determining the overall
isotope fractionation: 1. the equilibrium isotope fractionation of the liquid to
vapor phase transition, 2. a fractionation resulting from the diffusion across the
air boundary layer, and 3. the back flux of the atmospheric moisture.
March 16, 1996 8:49 Annual Reviews GAT-8 AR10-08
The most useful model for the isotope fractionation during evaporation,
which has stood the test of time, is that of Craig & Gordon (1965). The
model is based on the Langmuir linear-resistance model for evaporation and
is schematically shown in Figure 3. The following assumptions underlie this
model:
(a) equilibrium conditions at the air/water interface, so that the relative humidity
is h = 1 and RV = α ∗ RL ;
(b) a constant vertical flux, i.e. no divergence or convergence in the air column;
and
The appropriate flux equations for the water substance (E) and isotopic molecul-
es (E i , either for HDO or H2 18 O) are then
E = (1 − h)/ρ; ρ = ρM + ρT
and
234 GAT
236 GAT
isotopes and assuming marine air in equilibrium with the marine precipitation
at a temperature of 20◦ C (with the humidity h and the factor θ as open param-
eters) yield very reasonable values of h = 0.743 and θ = 0.48. This value of
θ ∼ 0.5 should be noted, as one encounters this value repeatedly over large
evaporative systems (Gat et al 1995) as noted above.
As the marine air then moves over the coast and into the continents, the differ-
ent marine air parcels appear to mix and homogenize, resulting in precipitation
that is closely aligned along the so-called meteoric water lines (MWL). These
are the lines in δ-space whose formula is δ(D) = 8δ(18 O) + d, where d has been
named the “deuterium excess” parameter by Dansgaard (1964); obviously the
GMWL (Craig 1961b) is one such meteoric water line with d = 10‰. The
further removed from the vapor source, the more depleted the isotopic values
of the precipitation on each of these lines. Dansgaard (1964) recognized four
parameters that determine this depletion in the isotopic values, including an
altitude effect, a distance from the coast, and a latitude effect. All of these
are basically related to the wringing out of moisture from the atmosphere as a
result of cooling of the air mass, and indeed the correlation with the temperature
appears as the overriding factor (Yurtsever 1975). Only the fourth parameter
described by Dansgaard, the “amount effect,” has a more complicated structure
(Gat 1980). Figure 7 shows the worldwide distribution of the oxygen-18 values
in annual precipitation.
To explain these findings of terms of the Rayleigh equations described above,
the early concepts on the evolution of the isotopic composition of “meteoric”
waters envisaged an air mass (with water vapor mass N and a given isotopic
composition Rv ) from which precipitation is derived by equilibrium condensa-
tion, so that at any time: Rp = α ∗ Rv . The system then follows a “Rayleigh
law,” according to Equation 1b, with the fractionation factor of α ∗ . In integrated
form λ = λ0 + ε∗ ln f , with λ ≡ lnR; (α ∗ − 1) ≡ ε∗ and f is the fraction of
remaining water vapor in the system ( f = N /N0 ).
This equation establishes a linear relationship in λD -λ18 space under isother-
mal conditions, i.e. when α ∗ remains fixed throughout the process. The slopes
of such Rayleigh lines in λ-space vary from the value of Sλ = εD∗ /ε18 ∗
= 8.22
◦ ◦
to a value of Sλ = 9.6 over the temperature range of 30 C to 0 C, according
the fractionation factors as given by Majoube (1971).
The λ- and δ(‰)-space are related by the equations:
µ ¶
R δ
= 1+ 3
Rstd 10
and
dδ/103
dλ = d[ln(1 + δ/103 )] = .
(1 + δ/103 )
March 16, 1996 8:49 Annual Reviews GAT-8 AR10-08
238 GAT
(a)
Figure 5 (a) The marine-atmosphere isotope model of Craig & Gordon (1965). The first of the δ
values refers to δ(18 O), while the value in parenthesis refers to the value of δ(D).
(b)
Figure 5 (b) Adaptation of the Craig-Gordon marine model to account for the terrestrial part of
the hydrologic cycle. ∞ indicates isotopic equilibrium between the vapor and precipitation fluxes.
March 16, 1996 8:49 Annual Reviews GAT-8 AR10-08
240 GAT
Figure 6 Monthly precipitation values at selected marine stations, based on data of the IAEA
network on a δ(D) vs δ(18 O) diagram (Rosanski et al 1992).
March 16, 1996
8:49
Annual Reviews
GAT-8
AR10-08
Figure 7 Distribution of the annual mean δ(18 O) value in precipitation, based on the IAEA network data. (From Yurtsever & Gat 1981.)
March 16, 1996 8:49 Annual Reviews GAT-8 AR10-08
242 GAT
(1964) and Yurtsever (1975), or more precisely with the temperature at the cloud
base (Rindsberger & Magaritz 1983). This simple scenario, which generally
fits the data rather well, is modified under exceptional circumstances, namely,
• in the case where snow or hail reaches the ground; the isotopic exchange
then does not occur, with the result that the precipitation is more depleted
than in the equilibrium situation. Often, in addition, the solid precipitation
forms show higher d-values due to nonequilibrium condensation during the
growth of ice particles (Jouzel & Merlivat 1984).
• in the case of precipitation from strongly convective systems [thunder clouds,
cold fronts, and tropical clouds associated with the ITCZ (intertropical con-
vergence zone)] which are characterized by strong local downdrafts. As a
result the raindrops do not interact with an averaged sample of the ambient
air but only with a portion of the in-cloud air.
In the two cases cited above, the isotopic value in the precipitation is more
depleted than the true equilibrium precipitation. These situations are then
less effective in isotopic fractionation during rainout than the Rayleigh pro-
cess proper, moderating the extent of isotopic depletion as a function of the
rainout.
An additional process that can affect the isotopic composition of precipitation
and of the atmospheric moisture is the evaporation from falling rain droplets
beneath the cloud base or from surface waters. Evaporation from the falling
rain results in the enrichment of the heavy isotopic species in the remnant drop
along the so-called evaporation line; since these usually have a slope of less than
8 in δ(D) vs δ(18 O) space, the resulting precipitation shows a smaller d-value
than at the cloud base. On the other hand, as shown in Figure 8, the evaporated
moisture’s isotope composition is characterized by larger d-excess value, so that
the precipitation derived from an air mass into which the reevaporated moisture
is admixed is also characterized by a large d-excess (Dansgaard 1964). A
similar effect occurs with the evaporated flux from surface waters (Goodfriend
et al 1989, Gat et al 1994).
The d-excess parameter has been shown to be a diagnostic tool for measuring
the contribution of evaporated moisture to the downwind atmosphere (Gat et al
1994). The scaling factor for judging a change in the d-value of the perturbed
atmosphere is given by
(dL − da ) ¡ ¢
(dE − da ) = + CD(D) − CD(18 O) × θ × n,
1−h
with subscripts E, a, and L standing for the evaporation flux, the ambient mois-
ture, and the liquid, respectively.
March 16, 1996 8:49 Annual Reviews GAT-8 AR10-08
244 GAT
Figure 8 Schematics of the addition of evaporated moisture from surface water into the ambient
atmosphere on the δ-diagram. δp = δ value of precipitation; δw = δ value of the residual water
after evaporation; δE = the evaporation flux; δa and δa0 are the atmospheric moisture before and
after mixing with δE .
and air-sea interaction conditions. This affects the d-value primarily. Further-
more, different seasons have different degrees of rainout as dictated by the
temperature along the air mass trajectory (which fixes the position along the
“Rayleigh” line). In addition, the deviation from the Rayleigh relationship by
either evaporative enrichment from falling raindrops, the admixture of reevapo-
rated moisture, or the effects of snow and hail are also different for the different
seasons.
A typical seasonal march of precipitation values is shown in Figure 9, based
on data provided by the IAEA precipitation network (IAEA 1992). Both the
heavy isotope content and the d-parameter vary over a yearly cycle, with the
d-values for the winter precipitation usually somewhat higher. Evidently, the
Figure 9 Seasonal change of δ(18 O) and d for selected continental stations, based on IAEA
network data (IAEA 1992).
March 16, 1996 8:49 Annual Reviews GAT-8 AR10-08
246 GAT
effect of high d-values in snow (Jouzel & Merlivat 1984) and the reduced
degree of evaporation from falling rain droplets in winter when compared to
the summer conditions usually override the effect of higher humidity at the
oceanic source areas, which would result in the opposite effect (Merlivat &
Jouzel 1979).
There are still insufficient data to fully characterize shorter term isotopic
changes during a storm or in between individual precipitation events. It ap-
pears that such variations span a wide range of δ-values, up to ± 10‰ in δ 18 O
at times. Some early studies by Epstein (1956), Bleeker et al (1966), and Mat-
suo & Friedman (1967) indicated differences between rain produced by warm
front uplifting or cold fronts and associated thunder clouds. These studies
also showed that the beginning part of most rainshowers are more enriched in
the heavy isotopes due to partial evaporation during the fall of rain droplets
through the air column. Later studies (Leguy et al 1983, Gedzelman et al 1989,
Rindsberger et al 1990, Pionke & Dewalle 1992, McDonnel et al 1990) make
it appear that these differences reflect mainly the source of moisture and its
rainout history, and only to a lesser extent the local rain intensity (EM Adar,
unpublished data from the Negev). However, a notable exception to this rule is
given by very strong tropical rains, associated with the ITCZ and its towering
clouds, when precipitation with extremely depleted isotopic values are found
at the peak of the downpour (Matsui et al 1983).
It is also of interest that in a continental setting the average air moisture and
the daily precipitation are close to isotopic equilibrium with each another (Craig
& Horibe 1967, Jacob & Sonntag 1991). This is not strictly true in a coastal
setting or in regions of climate transition (Matsui et al 1983).
4. PRECIPITATION-INFILTRATION-RUNOFF
RELATIONSHIPS OVER LAND
4.1 Surface Interception of Precipitation
As rain falls on the ground, a number of processes take place that modify the
isotope composition of the incoming precipitation. Details of these processes
depend on the nature of the terrain and on the characteristics of the rain, such
as its amount and duration, intensity, and intermittency.
As shown in Figure 10 [which is an adaptation of an early model of Gat &
Tzur (1967)], part of the incoming precipitation is intercepted on the canopy
of the vegetation cover and in part lost by evaporation. Up to 35% of the
precipitation is thus lost in tropical forests (Molion 1987) and 14.2 and 20.3%,
respectively, on deciduous and coniferous trees in the Appalachian Mountains
in the USA (Kendall 1993). While it evaporates, the water on the canopy,
March 16, 1996 8:49 Annual Reviews GAT-8 AR10-08
Figure 10 The interception reservoirs during the precipitation to runoff transition, based on Gat
& Tzur (1967).
248 GAT
intense parts of the rainshower are preferentially lost. Since these usually
constitute the isotopically most enriched part of the rain [corresponding to the
amount effect of Dansgaard (1964)], this selection process introduces a negative
bias that cancels the signature of the evaporation process.
The isotopic change on an interception volume V —where aV is the first
rainshower, f is the fraction of V remaining after evaporative water loss, and
bV is the follow-up shower—that results in throughflow when b > (1 − f ) can
be modeled with some simplifying assumptions. The isotopic composition of
the throughfall is then
R
f δA + (b − 1)δB f ε 0 ( f )d ln f
δTh = + , (8)
(b + f − 1) (b + f − 1)
where
dδ f h(δ f − δatm ) − ε
= = ε 0 ( f ), (9)
d ln f (1 − h)
and δA and δB are the isotopic composition of the first and second rain events,
respectively.
Whereas the first term is simply the weighted mean of the contribution of
the first and second rainy spell, it is the second term that expresses the change
in isotope composition due to evaporation. As was shown in a similar model
applied to evaporation from the soil column (Gat 1987), this term has a max-
imum for f ∼ 0.5 and vanishes as f approaches either 0 or 1. Although the
effect of evaporation from the canopy is thus difficult to quantify based on the
local recharge flux, it is reflected in the downwind atmosphere in the form of an
increase of the d-excess parameter of the atmospheric moisture. Gat & Matsui
(1991) thus interpreted an increase of close to 3‰ in the value of d over the
Amazon Basin as the cumulative effect of reevaporation from the canopy of
the rainforest. The transpiration flux, while larger than that of the interception
flux, is not expected to produce a change in d.
In the semi-arid zone the plant cover is less developed. Here the direct
evaporation from temporary surface water impoundments that appear following
stronger rains are expected to produce the largest isotopic signal. Gat & Tzur
(1967) calculated the change between the isotopic composition of infiltrating
water from such surface impoundments and the precipitation. For the climatic
conditions of Israel the enrichment is less than 1‰ [in δ(18 O)] on soils with
infiltration capacity of < 2 mm/hr in winter and ∼ 2‰ in summer. On heavy
soils, higher distortions of up to 5‰ are expected. The effect is minimal on
sandy soils where one does not encounter free surface waters. The effect of
evaporation from within the soil column, and especially from well-ventilated
sand, is discussed below.
March 16, 1996 8:49 Annual Reviews GAT-8 AR10-08
250 GAT
a tool for studying mixing in the lake. It is also a useful tracer for detecting the
addition of lake waters to adjacent groundwaters (Krabbenhoft et al 1994).
Under hydrologic steady-state conditions the isotopic buildup in a lake is
given by
(δa − δin + ε/ h)
1δ = δL,SS − δin = h i, (10)
1 + FEin · (1−h)
h
where Fin and E are the influx and evaporation terms, respectively. Evidently,
changes in either δin , δa , the humidity (h), or the hydrologic balance of the
lake are determining parameters. Higher enrichments up to a limiting value of
δL = δa + ε/ h are possible for situations such as a string of lakes, i.e. when a
number of evaporative elements act in series (Gat & Bowser 1991).
The subject of lakes, freshwater and saline ones, has been extensively re-
viewed over the years. Gat (1995) summarizes the state of the art in this
respect.
4.4 Soilwaters
Waters infiltrating from the surface drain through the void spaces in the soil,
once the “field capacity,” i.e. the amount of water held against gravity on
the grains of the soil by adsorption and capillary forces, is filled up. In very
homogeneous soils one can recognize a piston flow pattern where the infiltrate of
each season pushes the previous one ahead, as visualized by the tritium content
of the moisture column (Dincer et al 1974, Gvirtzman & Magaritz 1990). In
heterogeneous soils and fractured matrixes the pattern is more complicated.
The holdup in the soil column and transport through it does not, in itself, affect
the isotopic composition of the infiltrating waters, except as far as mixing be-
tween moving and standing water parcels occurs. However, when evaporation
from within the topsoil occurs during dry intervals between rain events, this re-
sults in an enrichment of the heavy isotopes in the residual waters (Zimmerman
et al 1967). At some depth beneath the surface an evaporating front develops:
Above it water transport is predominantly in the gaseous phase through the air-
filled intergranular space; below it laminar flow and molecular diffusion in the
water-filled pore space dominate. An isotopic concentration profile develops,
representing a balance between the upward convective flux and the downward
diffusion of the evaporative signature (Barnes & Allison 1988). Isotopic en-
richments of 18 O in excess of 10‰ relative to the precipitation are found in
desert soils. Because the isotope enrichment at the evaporating front occurs by
diffusion through a stagnant air layer, Equation 6a with n = 1 applies. The
isotopic composition of the residual waters are characterized by their situation
on a low-slope evaporation line (Figure 11). These enriched waters are then
March 16, 1996 8:49 Annual Reviews GAT-8 AR10-08
Figure 11 Isotopic relationships during the groundwater formation in the arid and semiarid zone.
March 16, 1996 8:49 Annual Reviews GAT-8 AR10-08
252 GAT
in the veins of the leaves [amounting at times up to 30% of the total leafwater
volume (Allison et al 1985)]. In the case of cotton plants, Yakir et al (1990), on
the other hand, postulated that part of the leafwater (in the mesophyllic cells)
lags behind the steady-state enrichment due to a slow water exchange with the
outer water sheath, which is assumed to be in momentary steady state relative
to the evaporation process during the day and to take up unfractionated source
water at night.
Under steady-state conditions obviously δE = δin , so that the transpiration
flux is unfractionated with respect to the soilwater taken up by the plants.
The enriched isotopic values of plant waters is being used to check the au-
thenticity of fruit products and distinguish them from those adulterated by water
(Bricout et al 1972).
5. GROUNDWATERS
5.1 Meteoric Waters
The isotopic composition of groundwaters that are recharged by direct infil-
tration in the temperature zone follows that of the incident precipitation rather
closely. The surface processes discussed above usually cause shifts in the iso-
tope composition of less than 1‰ in δ(18 O) (Gat & Tzur 1967). The isotope
composition of groundwaters are thus useful tools for identifying recharge sites
(utilizing geographic effects on the isotopic composition of rain, such as the
altitude effect), tracing mixing patterns and, in particular, for detecting the en-
croachment of lake waters (which are enriched in the heavy isotopes) (Payne
1970) of seawater and other extraneous water sources. In the arid zone, direct
infiltration can result in recharge only on sand. In other cases some surface
runoff is a necessary prerequisite for accumulating sufficient water depth to
enable occurrence of recharge. Both small-scale local runoff, feeding locally
important aquifer pockets in limestones and crystalline rocks, and large-scale
regional flood flows, which recharge large deep regional aquifers, are important
recharge mechanisms. Each of these recharge pathways imprints a distinguish-
able stable isotopic signature on the recharge waters (Figure 11).
In order to take full advantage of the possibilities of understanding groundwater
formation, the detailed isotope effects of the recharge and runoff processes need
to be established for different climate conditions and for each watershed. Such
studies are still few and incomplete. When available, the changing isotope com-
position in groundwater recharge or runoff could be used as a sensitive monitor
of climatic and anthropogenic changes in the watershed (Gat & Lister 1995).
Up to temperatures of about 60–80◦ C, the isotopic composition of groundwa-
ters is conserved in the aquifer, so that even waters recharged in the Pleistocene
March 16, 1996 8:49 Annual Reviews GAT-8 AR10-08
254 GAT
more than 10,000 years ago are still useful for climate reconstruction of these
past conditions based on the isotopic composition of paleowaters (Fontes 1981).
For example, the wetter climate of North Africa was established by interpret-
ing paleowaters in the Nubian sandstone aquifers (Sonntag et al 1976). By
virtue of the lower d-excess value of such waters (d ∼ 7‰), Merlivat & Jouzel
(1979) established a higher humidity over the Atlantic at that time. However,
the advantage of having a record for both oxygen and hydrogen isotopes in
paleowaters (whereas most other climate proxies only give the 18 O record) is
offset by the difficulty of reliably dating such waters.
model, there is then no a priori relation between the dissolved salt composition
and the isotopic composition of the water.
These initial solutions are then exposed to various processes. Three geo-
chemical processes can be considered. The first process is surface exposure
accompanied by evaporative water loss. According to Lloyd (1966), the iso-
tope composition tends in this case towards an enriched value while salinity
increases.
A second process is that of ultrafiltration by clay membranes. This process
results in a slight enrichment of both 18 O and D in the residual brine, correspond-
ing to a parallel increase in salinity with a preferential increase of multivalent
cations relative to sodium. The single stage isotope fractionation factors in
dilute NaCl solutions as a result of this process are 0.9993 for 18 O and 0.998 for
deuterium (Coplen & Hanshaw 1973). The fractionation effect is thus smaller
by one order of magnitude than the fractionation in the evaporation process.
The relationships between the increase of salinity and increase in 18 O content by
ultrafiltration for the solutions considered by Coplen & Hanshaw (under condi-
tions of Rayleigh fractionation) is given by the equation S = S0 exp(1δ 18 /0.7).
If one assumes that the process starts with seawater pushed through a clay mem-
brane filter until the dissolved salt increases by a factor of 4 to a concentration of
150 g/liter (which are concentrations typical of formation waters), then this rela-
tion predicts an enrichment of 18 O in the water by + 1‰. A parallel increase in
the Ca/Na ratio would be expected (White 1965). If additional freshwater then
flows through this system, a further increase of the Ca/Na ratio may result (form-
ing a CaCl2 type water), accompanied, however, by a possible slight decrease
in the concentration of total dissolved solids. The 18 O content of the residual
brine approaches in the steady-state limit an enrichment of 0.7‰ relative to the
meteoric water (this being the single stage fractionation factor for this process).
A third process that can affect the isotopic composition of formation waters is
interaction with minerals in the formation, especially at elevated temperatures.
In this case the changes in the chemical and isotopic compositions depend on
the nature of the minerals. One extreme case occurs when the fluid is in contact
with limestones: The 18 O content of the fluid increases by exchange, without
significant change of the salinity. Another extreme occurs in the presence of
anhydrous evaporite deposits, which when they dissolve, cause the salinity
to increase without significantly changing the isotope composition of the fluid.
White et al (1973) have claimed that in the presence of clays deuterium exchange
can occur, resulting in a slight decrease of the δ(D) values.
The three processes mentioned result in increases of salinity and enrichment
of 18 O in the brine (Figure 12). Mixing with meteoric waters has the opposite
effect, i.e. a decrease in salinity and depletion in 18 O. The evidence of the
March 16, 1996 8:49 Annual Reviews GAT-8 AR10-08
256 GAT
hydrogen isotopes is more ambiguous, since both exchange with clays and
mixing with meteoric waters results in slightly depleted δ(D) values.
Figure 12 shows the changes in salinity and isotopic composition predicted
for the waters, as these evolve through any of the processes described above or
when mixing with meteoric waters occurs.
enabled one to use cores in glaciers to establish the most reliable isotope record
of precipitation as far back as the Pleistocene (e.g. Dansgaard et al 1969). The
establishment of the isotope record in glaciers constitutes one of the highlights
of isotope hydrology, which is not discussed further here in view of the many
good descriptions available.
The sublimation of the ice removes the material layer by layer. It could
be expected then that this process would not alter the isotopic composition of
the remaining snowpack, unlike the situation in an evaporating water body in
which the effect of surface fractionation is propagated into the residual liquid
by mixing. However, this simple model is not realized in practice, apparently
due to diffusion of vapor through the void spaces in the snow pack and also due
to some surface melting and percolation of meltwaters into the remaining snow.
As a result of such processes, which have been extensively explored (Moser &
Stichler 1974, Whillans & Grootes 1985, Friedman et al 1991, also the review of
Arnason 1981) meltwaters do not always show the same isotopic composition as
that of the snow; the remainder of the snow pack shows increasing enrichment of
the heavy isotopes and reduced values of the d-parameter. An extreme example
is reported by Claasen & Downey (1995) in the canopy-intercepted snowfall
on evergreens, where enrichments of up to 2‰ in δ(18 O) are found.
Another facet of solid-phase isotope hydrology is the constitution of the sea-
ice cover and of lake ice. Merlivat (1978) has shown that a marine glacier is
made up mostly of accumulated precipitation and that only the very bottom
layers are frozen seawater. On a lake in northern Wisconsin, on the other hand,
Bowser & Gat (1995) found the lower half of the ice cover to be ice grown in
equilibrium with the lake water, which, however, is mixed with unfractionated
lake water trapped in the ice during the freezing process. The upper 50% of the
ice core is evidently made up of lake water that overflows on top of the subsiding
ice layer, mixes with the accumulated snow on the surface, and is then refrozen.
7. APPLICATIONS TO PALEOCLIMATOLOGY
The usefulness of the changes in isotope composition in the hydrologic cycle
for hydrological studies is self-evident. The essentially conservative behavior
of the isotope in the subsurface, on the one hand, and the characteristic and
well-documented changes in the atmospheric section of the cycle and at the
surface-atmosphere interface, on the other hand, provide a reliable tracer (fin-
gerprint). This tracer has been utilized to determine the origin of water bodies,
to quantify the water balance of surface and subsurface reservoirs, and to trace
the mechanisms of salinization of water resources (Gat 1975).
The isotope record in climatic archives is increasingly called upon to provide
paleoclimatic information. This analysis is based on two premises: (a) that the
March 16, 1996 8:49 Annual Reviews GAT-8 AR10-08
258 GAT
isotope composition of the proxy material can be related to that of the water cycle
of past periods and (b) that from the latter one can evaluate the relevant climatic
or hydrologic parameters. Both of these premises need a critical appraisal.
As for the first of these premises, the situation is fairly straightforward where
the proxy material is the water substance itself, as in ice cores, paleowaters, or
interstitial waters in sediment cores. However, the constancy of the isotopic
record has to be addressed, whether with regard to diffusive mixing in the
column, the interaction with the rock matrix or sedimentary material (affecting
the oxygen isotopes primarily), or the effect of biogeochemical interactions that
occur in porewaters, which affect mainly the hydrogen isotopes. The porewater
system cannot be regarded as a closed one for more than a few thousand years
at best because the older isotopic signals are erased by porewater diffusion (GS
Lister, unpublished data).
In the case of proxy materials such as lacustrine carbonates, shells of land-
snails (Goodfriend et al 1989), or plant materials (cellulose), the major issue
is the relationship of the isotopic composition in the proxy to the environ-
mental waters. Lake carbonate deposits are related to the lake waters through
a temperature-dependent fractionation factor (often compounded by an addi-
tional “vital effect,” i.e. a biochemical effect). The lake waters, in turn, are
related to the inflowing waters isotope composition and to the degree of isotopic
enrichment of the lake waters.
Thus both hydrological and climatic factors are involved, as well as ecological
ones, as discussed by Talbot (1990) and Gat & Lister (1995).
Plant material, on the other hand, is related to the hydrological cycle through
the soilwater taken up in the roots. The isotope composition is, however,
modified by hydrological and physiological processes in the plant, such as
water stress, ambient humidity, and also the timing of stomata openings.
In all these cases the isotope composition of the precipitation is thus modified
by factors that are climatically and environmentally controlled and that need to
be recognized in great detail under a variety of climatic conditions.
The second premise mentioned above—which concerns the interpretation
of changes in the isotope composition of precipitation with time in terms of
changing climate parameters, temperature in particular—is based on the strong
correlation between the isotopic composition of precipitation and the ambient
temperature in the present-day water cycle. Various authors have attempted
to impose this relationship on the paleo-record. This obviously could be an
illusive procedure when one considers that the sources and sinks of water as
well as details of the global circulation undergo important changes over the
periods concerned. The approach of modelers such as Joussaume & Jouzel
(1993), who attempt to derive global isotope maps for past climate conditions
March 16, 1996 8:49 Annual Reviews GAT-8 AR10-08
as reference points for the local climatic record, is a more promising approach.
This effort, however, requires a rather detailed understanding of the bound-
ary conditions both over the ocean and at the plant (soil)-atmosphere inter-
face.
It is interesting to note that the interest in the paleoclimatic applications of
isotope hydrology has been a stronger motivation for the study of stable isotopes
in the hydrologic cycle (yielding much useful hydrological information as a by-
product) than that provided by the hydrological sciences directly.
Any Annual Review chapter, as well as any article cited in an Annual Review chapter,
may be purchased from the Annual Reviews Preprints and Reprints service.
1-800-347-8007; 415-259-5017; email: [email protected]
Literature Cited
Adar EM, Gev I, Lip J, Yakir D, Gat JR. 1995. in nature. Summer Course on Nuclear Ge-
Utilization of oxygen-18 and deuterium in ology, Varenna 1960. Lab. Geol. Nucl. Pisa.
stream flow for the identification of transpira- 129 pp.
tion sources: soil water versus groundwater Bolin B. 1959. On the use of tritium as a tracer
in sand dune terrain. In Application of Trac- for water in nature. In Proc. 2nd Conf. on the
ers in Arid Zone Hydrology, ed. E Adar, C Peaceful Uses of Atomic Energy, 18:336–44.
Leibundgut. Int. Assoc. Sci. Hydrol. Publ. Geneva: United Nations
232:329–38 Bowser C, Gat JR. 1995. On the process of
Allison GB, Gat JR, Leaney FWJ. 1985. The re- lake ice formation. In Isotopes in Water Re-
lationship between deuterium and oxygen-18 sources Management. Vienna: Int. At. En-
values in leaf water. Isotope Geosci. 58:145– ergy Agency. In press
56 Bricout J, Fontes JCh, Merlivat L. 1972. Sur la
Arnason B. 1969. Equilibrium constant for the composition en isotopes stables de l’eau de
fractionation of deuterium between ice and jus d’oranges. C. R. Acad. Sci. Paris Ser. D
water. J. Phys. Chem. 73:3491–94 274:1803–6
Arnason B. 1981. Snow and ice hydrology. In Brutsaert W. 1965. A model for evaporation as
Stable Isotope Hydrology, ed. JR Gat, R Gon- a molecular diffusion process into a turbulent
fiantini. IAEA Tech. Rep. Ser. 210, Chap. 7, atmosphere. J. Geophys. Res. 70:5017–24
pp. 143–75 Chow VT. 1964. Handbook of Applied Hydrol-
Baertschi P. 1976. Absolute 18 O content of Stan- ogy. New York: McGraw Hill
dard Mean Ocean Water. Earth Planet. Sci. Claasen HC, Downey JS. 1995. A model for
Lett. 31:341–44 deuterium and oxygen isotope changes dur-
Barnes CJ, Allison GB. 1988. Tracing of wa- ing evergreen interception. Water Resources
ter movement in the unsaturated zone using Res. 31:601–18
stable isotopes of hydrogen and oxygen. J. Clayton RN, Friedman I, Graf DL, Mayeda T,
Hydrol. 100:143–76 Meents WF, Shimp NF. 1966. The origin of
Bigeleisen J, Lee M, Mandel F. 1973. Equilib- saline formation waters—I. Isotopic compo-
rium isotope effects. Annu. Rev. Phys. Chem. sition. J. Geophys. Res. 71:3869–82
24:407–40 Coplen TB, Hanshaw B. 1973. Ultrafiltration by
Bigeleisen J, Mayer M. 1947. Calculation of a compacted clay membrane. I. Oxygen and
equilibrium constants for isotopic exchange hydrogen isotopic fractionation. Geochim.
reactions. J. Chem. Phys. 15:261–67 Cosmochim. Acta 37:2295–310
Bleeker W, Dansgaard W, Lablans WN. 1966. Craig H. 1961a. Isotopic variations in meteoric
Some remarks on simultaneous measure- waters. Science 133:1702–8
ments of particulate contaminants including Craig H. 1961b. Standards for reporting concen-
radioactivity and isotopic composition of pre- trations of deuterium and oxygen-18 in natu-
cipitation. Tellus 18:773–85 ral waters. Science 133:1833–34
Boato G. 1961. Isotope fractionation processes Craig H, Gordon LI. 1965. Deuterium and
March 16, 1996 8:49 Annual Reviews GAT-8 AR10-08
260 GAT
oxygen-18 variations in the ocean and marine Friedman I. 1953. Deuterium content of natural
atmosphere. In Stable Isotopes in Oceano- water and other substances. Geochim. Cos-
graphic Studies and Paleo-Temperatures, ed. mochim. Acta 4:89–103
E Tongiorgi, pp. 9–130. Pisa: Lab. Geol. Friedman I, Benson C, Gleason J. 1991. Iso-
Nucl. topic changes during snow metamorphism.
Craig H, Horibe Y. 1967. Isotope characteris- In Stable Isotope Geochemistry. Spec. Publ.
tics of marine and continental water vapour. Geochem. Soc. 3:211–21
Trans. Am. Geophys. Union 48:135–36 Friedman I, Machta L, Soller R. 1962. Water
Dansgaard W. 1953. The abundance of 18 O in vapour exchange between a water droplet and
atmospheric water and water vapour. Tellus its environment. J. Geophys. Res. 67:2761–66
5:461–69 Fritz P, Cherry JA, Weyer KV, Sklash MG. 1976.
Dansgaard W. 1964. Stable isotopes in precipi- Storm runoff analysis using environmental
tation. Tellus 16:436–68 isotopes and major ions. In Interpretation of
Dansgaard W, Johnsen SJ, Moller J, Langway Environmental Isotopes and Hydrochemical
CC Jr. 1969. One thousand centuries of cli- Data in Groundwater Hydrology, pp. 111–30.
matic record from Camp Century on the Vienna: Int. At. Energy Agency
Greenland Ice sheet. Science 166:377–81 Fritz P, Drimmie RJ, Frappe SK, O’Shea K.
Dawson TE, Ehleringer JR. 1991. Streamside 1987. The isotopic composition of precipi-
trees that do not use stream water. Nature tation and groundwater in Canada. In Isotope
350:335–37 Techniques in Water Resources Development,
Dewalle DR, Swistock BE. 1994. Differences in pp. 539–50. Vienna: Int. At. Energy Agency
oxygen-18 content of throughfall and rainfall Gat JR. 1975. Elucidating salination mechanism
in hardwood and coniferous forests. Hydrol. by stable isotope tracing of water sources. In
Proc. 8:75–82 Brackish Water as a Factor in Development,
Dewit JC, VanderStraaten CM, Mook WG. ed. A Issar, pp. 15–23. Beersheva, Israel: Ben
1980. Determination of the absolute hydro- Gurion Univ.
gen isotopic ratio of V-SMOW and SLAP. Gat JR. 1980. The isotopes of hydrogen and oxy-
Geostandards Newslett. 4:33–36 gen in precipitation. In Handbook of Environ-
Dincer T, Al-Mughrin A, Zimmermann U. 1974. mental Isotope Geochemistry, ed. P Fritz, JCh
Study of the infiltration and recharge through Fontes, 1:21–47. Amsterdam/Oxford/New
the sand dunes in arid zones with special refer- York: Elsevier
ence to the stable isotopes and thermonuclear Gat JR. 1981. Paleoclimate conditions in the
tritium. J. Hydrol. 23:79–109 Levant as revealed by the isotopic compo-
Ehhalt D, Roether W, Vogel JC. 1963. A survey sition of paleowaters. Israel Meteorol. Res.
of natural isotopes of water in South Africa. Pap. 3:13–28
In Radioisotopes in Hydrology, pp. 407–15. Gat JR. 1987. Variability (in time) of the isotopic
Vienna: Int. At. Energy Agency composition of precipitation: consequences
Epstein S. 1956. Variations in the 18 O/16 O ra- regarding the isotope composition of hydro-
tios of freshwater and ice. Natl. Acad. Sci. logic systems. In Isotope Techniques in Water
Nucl. Sci. Ser. Rep. 19:20–28 Resources Development, pp. 551–63. Vienna:
Eriksson E, Khunakassem V. 1969. Chloride Int. At. Energy Agency
concentration in groundwater, recharge rate Gat JR. 1995. Stable isotopes of fresh and saline
and rate of deposition of chloride in the Israel lakes. In Physics and Chemistry of Lakes, ed.
Coastal Plain. J. Hydrol. 7:178–97 A Lerman, D Imboden, J Gat, pp. 139–66.
Facy L, Merlivat L, Nief G, Roth E. 1963. New York: Springer-Verlag
The study of the formation of a hailstone by Gat JR, Bowser C. 1991. The heavy isotope en-
means of isotopic analysis. J. Geophys. Res. richment of water in coupled evaporative sys-
68:3841–48 tems. In Stable Isotope Geochemistry, ed. HP
Fleisher E, Goldberg M, Gat JR, Magaritz M. Taylor, JR O’Neil, IR Kaplan. Geochem. Soc.
1977. Isotopic composition of formation wa- Spec. Publ. 3:159–68
ters from deep drillings in southern Israel. Gat JR, Bowser C, Kendall C. 1994. The contri-
Geochim. Cosmochim. Acta 41:511–25 bution of evaporation from the Great Lakes to
Fontes JCh. 1981. Paleowaters. In Stable Isotope the continental atmosphere: estimate based
Hydrology, ed. JR Gat, R Gonfiantini. IAEA on stable isotope data. Geophys. Res. Lett.
Tech. Rep. Ser. 210, Chap. 12, pp. 273–98 21:557–60
Fontes JCh, Yousfi M, Allison GB. 1986. Esti- Gat JR, Carmi I. 1970. Evolution of the iso-
mation of long term diffuse groundwater dis- topic composition of atmospheric waters in
charge in the northern Sahara using stable iso- the Mediterranean Sea area. J. Geophys. Res.
tope profiles in soil water. J. Hydrol. 86:315– 75:3039–48
27 Gat JR, Lister GS. 1995. The “catchment effect”
March 16, 1996 8:49 Annual Reviews GAT-8 AR10-08
on the isotopic composition of lake waters; its Krabbenhoft DP, Bowser CJ, Kendall C, Gat JR.
importance in palaeolimnological interpreta- 1994. Use of oxygen-18 and deuterium to as-
tions. In Proc. Eur. Sci. Found. Workshop. In sess the hydrology of groundwater-lake sys-
press tems. In Environmental Chemistry of Lakes
Gat JR, Matsui E. 1991. Atmospheric water and Reservoirs, ed. LA Baker, ACS Adv.
balance in the Amazon basin: an isotopic Chem. Ser. 237:67–90
evapotranspiration model. J. Geophys. Res. Leguy C, Rindsberger M, Zangwil A, Issar
96:13179–88 A, Gat JR. 1983. The relation between the
Gat JR, Shemesh A, Tziperman E, Hecht A, oxygen-18 and deuterium contents of rain-
Georgopoulis D, Ozden B. 1995. The stable water in the Negev Desert and air mass tra-
isotope composition of water of the eastern jectories. Isotope Geosci. 1:205–18
Mediterranean Sea. J. Geophys. Res. In press Leopoldo RR. 1981. Aspetos hidrologicos di
Gat JR, Tzur Y. 1967. Modification of the iso- florista amazonica denga na reglao de Man-
topic composition of rainwater by processes aus. PhD thesis. Univ. Natl. Estado San Paulo,
which occur before groundwater recharge. In Botucatu, San Paulo, Brazil
Isotopes in Hydrology, pp. 49–60. Vienna: Levin M, Gat JR, Issar A. 1980. Precipitation,
Int. At. Energy Agency flood and groundwaters of the Negev high-
Gedzelman SD, Rosenbaum JH, Lawrence JR. lands: an isotopic study of desert hydrology.
1989. The megalopitan snowstorm of 11–12 In Arid Zone Hydrology: Investigation with
February 1983: isotopic composition of the Isotope Techniques, pp. 3–22. Vienna: Int.
snow. J. Atmos. Sci. 46:1637–49 At. Energy Agency
Gonfiantini R, Gratziu S, Tongiorgi E. 1965. Lloyd RM. 1966. Oxygen isotope enrichment
Oxygen isotope composition of water in of sea-water by evaporation. Geochim. Cos-
leaves. In Isotopes and Radiation in Soil- mochim. Acta 30:801–14
Plant Nutrition Studies, pp. 405–10. Vienna: Macklin WC, Merlivat L, Stevenson CM. 1970.
Int. At. Energy Agency The analyses of a hailstone. Q. J. R. Meteorol.
Goodfriend G, Magaritz M, Gat JR. 1989. Stable Soc. 96:472–86
isotope composition of land snail body water Majoube M. 1971. Fractionment en oxygene-
and its relation to environmental water and 18 et en deuterium entré l’eau at sa vapeur. J.
shell carbonate. Geochim. Cosmochim. Acta Chim. Phys. 10:1423–36
53:3208–21 Matsui E, Salati E, Ribeiro M, Reis CM, Tan-
Gvirtzman H, Magaritz M. 1990. Water and an- credi A, Gat JR. 1983. Precipitation in the
ion transport of the unsaturated zone traced by Central Amazon Basin: the isotopic com-
environmental tritium. In Inorganic Contami- position of rain and atmospheric moisture at
nants in the Vadose Zone, ed. B Bar Yosef, NJ Belem and Manaus. Acta Amazonica 13:307–
Barrow, J Goldschmidt. Ecol. Stud. 74:190– 69
98. Berlin/Heidelberg/New York: Springer- Matsuo S, Friedman I. 1967. Deuterium content
Verlag of fractionally collected rainwater. J. Geo-
Hitchon B, Friedman I. 1969. Geochemistry phys. Res. 72:6374–76
and origin of formation waters in the western McDonnel JJ, Bonell M, Stewart MK, Pearce
Canada sedimentary basin. Geochim. Cos- AJ. 1990. Deuterium variations in storm rain-
mochim. Acta 33:1321–49 fall: Implications for stream hydrograph sep-
Jacob H, Sonntag Ch. 1991. An 8-year record aration. Water Resources Res. 26:455–58
of the seasonal variation of 2 H and 18 O in at- McKinney CR, McCrea JM, Epstein S, Allen
mospheric water vapour and precipitation at HA, Urey HC. 1950. Improvements in mass-
Heidelberg, Germany. Tellus 43B:291–300 spectrometers for the measurement of small
Joussaume S, Jouzel J. 1993. Paleoclimatic trac- differences in isotope abundance ratios. Rev.
ers: an investigation using an atmospheric Sci. Instrum. 21:724–30
general circulation model under ice age con- Melander L. 1960. Isotope Effects on Reaction
ditions. J. Geophys. Res. 98:2807–30 Rates. New York: Roland
Jouzel J, Merlivat L. 1984. Deuterium and Merlivat L. 1978. Molecular diffusivities of
oxygen-18 in precipitation: modelling of the H2 16 O, HD16 O and H2 18 O in gases. J. Chim.
isotopic effects during snow formation. J. Phys. 69:2864–71
Geophys. Res. 89:11749–57 Merlivat L, Contiac M. 1975. Study of mass
Jouzel J, Merlivat L, Roth E. 1975. Isotopic transfer at the air-water interface by an iso-
study of hail. J. Geophys. Res. 80:5015–30 topic method. J. Geophys. Res. 80:3455–64
Kendall C. 1993. Impact of isotopic heterogene- Merlivat L, Jouzel J. 1979. Global climatic in-
ity in shallow systems on modeling of storm- terpretation of the deuterium-oxygen 18 re-
flow generation. PhD thesis. Dept. Geology, lationship for precipitation. J. Geophys. Res.
Univ. MD, College Park. 270 pp. 84:5029–33
March 16, 1996 8:49 Annual Reviews GAT-8 AR10-08
262 GAT
Molion LCB. 1987. Micro-meteorology of an Sklash MG, Farvolden RN, Fritz P. 1976. A con-
Amazonian rainforest. In The Geophysiology ceptual model of watershed response to rain-
of Amazonia, ed. RE Dickinson. New York: fall, developed through the use of oxygen-18
Wiley. 285 pp. as a natural tracer. Can. J. Earth Sci. 13:271–
Mook WG, Groeneveld DJ, Brouwn AE, van 83
Ganswijk AJ. 1974. Analysis of a runoff hy- Sonntag C, Neureuther P, Kalinke Ch, Mu-
drograph by means of natural 18 O. In Isotope nich KO. 1976. Zur Paleoklimatik der Sahara:
Techniques in Groundwater Hydrology 1974, Kontinental Effect in D und 18 O Gehalt Plu-
1:145–55. Vienna: Int. At. Energy Agency vialer Sahara Wasser. Naturwissenschaften
Moser H, Stichler W. 1974. Deuterium and 63:479
oxygen-18 contents as an index of the prop- Talbot MR. 1990. A review of the paleohydro-
erties of snow covers. In Proc. of the Grindle- logical interpretation of carbon and oxygen
wald Symp. Int. Assoc. Sci. Hydrol. Publ. isotopic ratios in primary lacustrine carbon-
114:122–35 ates. Isotope Geosci. 11:261–80
Nier AO. 1947. A mass-spectrometer for isotope Taylor AB. 1972. The vertical variations of the
and gas analysis. Rev. Sci. Instrum. 18:398– isotopic concentrations of tropospheric water
404 vapour over continental Europe and their re-
Panichi C, Gonfiantini R. 1981. Geothermal lationship to tropospheric structure, NZ Inst.
waters. In Stable Isotope Hydrology, ed. JR Nucl. Sci. Rep. INS-R-107. 45 pp.
Gat, R Gonfiantini. IAEA Tech. Rep. Ser. 210, Urey HC. 1947. The thermodynamic proper-
Chap. 11, pp. 241–72 ties of isotopic substances. J. Chem. Soc.
Payne BR. 1970. Water balance of Lake Chala 1947:562–87
and its relation to groundwater from tritium Whillans IM, Grootes PM. 1985. Isotopic diffu-
and stable isotope data. J. Hydrol. 11:47–58 sion in cold snow and firn. J. Geophys. Res.
Pionke HB, Dewalle DR. 1992. Intra and inter- 90:3910–18
storm 18 O trends for selected rainstorms in White D, Barnes I, O’Neil J. 1973. Thermal
Pennsylvania. J. Hydrol. 138:131–43 and mineral waters of non-meteoric origin,
Rankama K. 1954. Isotope Geology. Oxford: California Coast range. Bull. Geol. Soc. Am.
Pergamon. 490 pp. 84:547–60
Rindsberger M, Jaffe M, Rahamim S, Gat JR. White DE. 1965. Saline waters of sedimentary
1990. Patterns of the isotopic composition of rocks. Am. Assoc. Petrol. Geol. Mem. 4:342–
precipitation in time and space: data from the 66
Israeli storm water collection program. Tellus Yakir D, DeNiro MJ, Gat JR. 1990. Natural deu-
42B:263–71 terium and oxygen-18 enrichment in leaf wa-
Rindsberger M, Magaritz M. 1983. The relation ter of cotton plants grown under wet and dry
between air-mass trajectories and the water conditions: evidence for water compartmen-
isotope composition of rain in the Mediter- tation and its dynamics. Plant, Cell, Environ.
ranean Sea area. Geophys. Res. Lett. 10:43– 13:49–56
46 Yakir D, Yechieli Y. 1995. Plant invasion of
Rozanski K, Araguas-Araguas L, Gonfiantini R. newly exposed hypersaline Dead Sea shores.
1993. Isotopic patterns in modern global pre- Nature 374:803–5
cipitation. In Climate Change in Continental Yurtsever Y. 1975. Worldwide survey of sta-
Isotopic Record, ed. PK Swart, KL Lohwan, ble isotopes in precipitation. Rep. Isotope
JA McKenzie, S Savin, Geophys. Monogr. 78, Hydrology Section. Vienna: Int. At. Energy
pp. 1–37. Washington, DC: Am. Geophys. Agency. 40 pp.
Union Yurtsever Y, Gat JR. 1981. Atmospheric waters.
Salati E, Dall’ollio A, Matsui E, Gat JR. In Stable Isotope Hydrology, ed. JR Gat, R
1979. Recycling of water in the Amazon Gonfiantini. IAEA Tech. Rep. Ser. 210, pp.
Basin, an isotopic study. Water Resources 103–42
Res. 15:1250–58 Zimmermann U. 1979. Determination by stable
Saxena RK. 1986. Estimation of canopy reser- isotopes of underground inflow and outflow
voir capacity and oxygen-18 fractionation in and evaporation of young artificial ground-
throughfall in a pine forest. Nordic Hydrol. water lakes. In Isotopes in Lake Studies, pp.
17:251–60 87–102. Vienna: Int. At. Energy Agency
Simpson E, Thornd DB, Friedman I. 1970. Dis- Zimmermann U, Ehhalt D, Munnich KO. 1967.
tinguishing seasonal recharge to groundwater Soil water movement and evapotranspiration:
by deuterium analysis in southern Arizona. In changes in the isotope composition of water.
World Water Balance, pp. 112–21. Int. Assoc. In Isotopes in Hydrology, pp. 567–84. Vienna:
Sci. Hydrol. Int. At. Energy Agency