You are on page 1of 121

I , ,

\, e '2...
Donorlo or q;¡,,;,
Developments in Petroleum Science, 1
Virgil Wink'e d/tr.'i)
, "~¡ 1,
''',,,~,¡.í~~,
?
,,> •• ,,'; •...•. ,".'

GEOCHEMISTRY OF OILFIELD
WATERS

A. GENE COLLINS

Bartlesville Energy Researcli Center


Bureau of Mines
United States Department of the Interior
Bartlesville, Oklahoma, U.S.A.

ELSEVIER SCIENTIFIC PUBLISHING COMPANY


Amsterdam Oxford New York 1975
ELSEVIER SCIENTIFIC PUBLISHING COMPANY
335 Jan van Galenstraat
P.O. Box 211, Amsterdam, The Netherlands

AMERICAN ELSEVIER PUBLISHING COMPANY, INC.


52 Vanderbilt Avenue
New York, New York 10017

Library of Congress Card Number: 73-89149

ISBN 0-444-41183-6

With 132 illustrations and 87 tables

Copyright © 1975 by Elsevier Scientific Publishing Company , Amsterdam


All rights reserved. No part oí this publication may be reproduced, stored in a retrieval
system, or transmitted, in any form or by any means, electronic, mechanical, photo-
copying, or otherwise without the prior written permission of the publisher, Elsevier
Scientific Publishing Company, Jan van Galenstraat 335, Amsterdam

Printed in The Netherlands


Chapter 1. INTRODUCTION

Petroleum, known to exist long before an oil well was drilled, f-irstfound
limited use as a medicine, lubricant, and waterproofing agent. The American
Indians knew of several oil and gas springs and gave this information to the
early American settlers. Early settlements were commonly located close to
salt licks which supplied salt to the population. Often these salt springs were
contaminated with petroleum, and many of the early efforts to acquire more
salt by digging wells were rewarded by finding unwanted increased amounts
of oil and gas associated with the saline waters. In the Appalachians, many
saline water springs occurred along the crests of anticlines (Rogers and
Rogers, 1843).
In 1855 it was found that distillation of petroleum produced a light oil
similar to coal oil, which was better than whale oil as an illuminant (Howell,
1934, p.2). This knowledge spurred the search for saline waters which con-
tained oi!. Colonel Edward Drake, uti!izing the methods of the salt pro-
ducers, drilled a well on Oil Creek, near Titusville, Pennsylvania, in 1859. He
struck oil at a depth of 21 m, and this first oil well produced about 35 barreIs
of oil per day (Dickey, 1959).
The early oil producers did not realize the significance of the oil and saline
waters occurring together. In fact, it was not until1938 that the existence of
interstitial water in oil reservoirs was generally recognized (Schilthuis, 1938).
Torrey (1966) was convinced as early as 1928 that dispersed interstitial
water existed in oil reservoirs, but his belief was rejected by his colleagues
because most of the producing oil wells did not produce any water upon
completion. Occurrences of mixtures of oil and gas with water were rec-
ognized by Griswold and Munn (1907), but they believed that there was a
definite separation of the oil and water, and that oil, gas, and water mixtures
did not occur in the sand before a well tapped the reservo ir.
It was not until 1928 that the first commerciallaboratory for the analysis
of rock cores was established (Torrey, 1966); the first core tested was from
the Bradford Third Sand (from the Bradford field, McKean County, Pennsyl-
vania). The percent saturation and percent porosity of this core were plotted
versus depth to construct a graphic representation of the oil and water
saturation. The soluble mineral salts that were extracted from the core led
Torrey to suspect that water was indigenous to the oil productive sand.
Shortly thereafter a test well was drilled near Custer City, Pennsylvania,
which encountered higher than average oil saturation in the lower part of the
Bradford Sand. This high oil saturation resuIted from the action of an un-
INTRODUCTION 3

tions proved that interstitial water was generally present in the oil sands.
Publications by Garrison (1935) and Schilthuis (1938) give detailed informa-
tion concerning the distribution of water and oil in porous rocks, and of the
origin and occurrence of "connate" water with information concerning the
relationship of water saturation to formation permeability.
The word "connate" was first used by Lane and Gordon (1908) to mean
interstitial water that was deposited with the sediments. The processes of
rock compaction and mineral diagenesis result in the expulsion of large
amounts of water from sediments and movement out of the deposit through
the more permeable rocks. It is therefore highly unlikely that the water
now in any pore is the same as that which was there when the particles that
surround it were deposited. White (1957) redefined connate water as "fossil"
water; it has been out of contact with the atmosphere for an appreciable part
of a geologic time periodo Connate water is thus distinguished frorn meteoric
water which has entered the rocks in geologically recent times, and frorn
juvenile water which has come frorn deep in the earth 's crust and has never
been in contact with the atmosphere.
Meanwhile petraleum engineers and geologists had learned that water s
associated with petroleum could be identified with regard to the reservo ir in
which they occurred by a knowledge of their chemical characteristics. Corn-
monly the water s from different strata differ considerably in their dissolved
chemical constituents, making the identification of a water from a particular
strata easy. However, in some areas the concentrations of dissolved con-
stituents in waters from different strata do not differ significantly, and the
identification of such waters is difficult or impossible.
The amount of water produced with the oil often increases as the amount
of oil praduced decreases. If this is edge water, nothing can be done about it.
If it is bottom water, the well can be plugged back. However, it often is
intrusive water from a shallow sand gainjng access to the well from a leaky
casing or faulty completion and this can be repaired.
Enormous quantities of water are produced with the oil in some fields,
and it is necessary to separate the oil frorn the water. Most of the oil can be
removed by settling. Often, however, an oil-in-water emulsion forms which is
very difficult to break. In such cases, the oil is heated and various surface-
active chemicals are added to induce separation.
In the early days, the water was dumped on the ground where it seeped
below the land surface. Until about 1930, the oilfield waters were disposed
into local drainage, frequently killing fish and even surface vegetation. After
1930, it became common practice to evaporate the water in earthen pits or
to inject it into the praducing sand or another deep aquifer. The primary
concern in such disposal practice is to remove all oil and basic sediment from
the waters before pumping them into injection wells, to preverÍt clogging of
the pore spaces in the formation receiving the waste water. Chemical corn-
patibility of waste water and host aquifer water must also be assured.
Waters produced with petroleum are growing in importance. In years past,
4 INTRODUCTION

these waters were considered waste and had to be disposed of in some


manner. Injection of these waters into the petroleum reservoir rock serves
three purposes: (1) it produces additional petroleum (secondary recovery);
(2) it utilizes a potential pollutant; and (3) in some areas it controls land
subsidence.
The volume of water produced with petroleum in the United States is very
large. In 1970, daily production was about 3.78 trillion liters of water with
about 1.51 trillion liters of oil. In older fields, the production is frequently
95% water and 5% petroleum.
Waterflooding in petroleum production is expanding rapidly, and in 1970
one-third to one-half of the production in the United States came from fields
into which water was injected. The volume of injected water has grown each
year. In many fields the volume of petroleum produced by secondary re-
covery by waterflooding is equal to the volume recovered by primary
methods.
To inject these waters into reservoir rocks, suspended solids and oil must
be removed from the waters to prevent plugging of the porous formations.
Water injection systems require separators, filters, and, in some areas,
deoxygenating equipment utilizing chernical and physical control methods to
minimize corrosion and plugging-in the injection system.
In waterflooding most petroleum reservoirs, the volume of produced
water is not sufficient to efficiently recover the additional petroleum. There-
fore, supplemental water must be added to the petroleum reservoir. The use
of waters from other sources requires that the blending of the produced
water with supplemental water must yield a chemically stable mixture so
that plugging solids will not be formed. For example, a produced water
containing considerable calcium should not be mixed with a water con-
taining considerable carbonate because calcium carbonate may precipitate
and prevent injection of the flood water. The design and successful operation
of a secondary recovery waterflood requires a thorough knowledge of the
composition of the waters used.
Chemical analyses of waters produced with oil are useful in oil production
problems, such as identifying the source of intrusive water, planning water-
flood and salt-water disposal projects, and treating to prevent corrosion
problems in primary and secondary recovery. Electrical well-log interpreta-
tion requires a knowledge of the dissolved solids concentration and composi-
tion of the interstitial water. Such information also is useful in correlation of
stratigraphic units and of the aquifers within these units, and in studies of
the movement of subsurface waters. It is impossible to understand the
processes that accumulate petroleum or other minerals without insight into
the nature of these waters.
REFERENCES 5

References
Dickey , P. A., 1959. The first oil well. J. Peto Technol., 11: 14-26.
Fettke, C.R., 1938. Bradford oil fie!d, Pennsylvania, and New York. Pa. Geol. Surv.,
Fourth Ser., Bull., M21: 1-454.
Garrison, A.D., 1935. Se!ective wetting of reservo ir rock and its relatíon to oí! produc-
tion. In: Drilling and Production Practice. American Petro!eum Institute, New York,
N.Y., pp.130-140.
Griswo!d, W.T. and Munn, M.J., 1907. Geo!ogy of oi! and gas fields in Steubenville,
Burgettstown and C!aysville Quadrang!es, Ohio, West Virginia and Pennsy!vania. U.S.
Geol. Surv. Bull., No.318, 196 pp.
Howell, J. V., 1934. Historica! development of the structural theory of accumulation of
oil and gas. In: W.E. Wrather and F.H. Lahee (Editors), Problems of Petroleum
Geology. American Association of Petroleum Geologists, Tulsa, Okla., pp.1-23.
Lane, A.C. and Gordon, W.C., 1908. Mine waters and their field assay. Bull. Geol. Soco
Am., 19:501-512.
Milis, R. van A., 1920. Experimental studies of subsurface relationships in oil and gas
fields. Econ. Geol., 15:398-421.
Munn, M.J., 1920. The anticlinal and hydraulic theories of oil and gas accumulation.
Econ. Geol., 4:509-529.
Rich, J.L., 1923. Further notes on the hydraulic theory of oil migration and accumula-
tion. Bull. Am. Assoc. Peto Geol., 7:213-225.
Rogers, W.B. and Rogers H.D., 1843. On the connection of thermal springs in Virginia
with antic1inal axes and faults. Am. Geol. Rep., pp.323-347.
Schilthuis, R.J., 1938. Connate water in oil and gas sands. In: Petroleum Development
and Techriology , AIME, pp.199-214.
Torrey , P.D., 1950. A review of secondary recovery of oil in the United States. In:
Secondary Recovery of Gil in the United States. American Petroleum Instítute, New
York, N.Y., pp.3-29.
Torrey, P.D., 1966. The discovery of interstitial water. Prod. Monthly , 30:8-12.
White, D.E., 1957. Magmatic, connate, and metamorphic water. Bull. Geol. Soco Am.,
68:1659-1682. .
Chapter 6. ORGANIC CONSTITUENTS IN SALINE WATERS

Water is a peculiar solvent and has been considered to possess at ambient


temperature a quasi-crystalline, open structure which will allow solute mole-
cules to fill the space between the lattice points (Eley, 1939). As a nonpolar
molecule dissolves in water at ambient temperature, the structure of the
water in its immediate vicinity becomes more crystalline, or a microscopic
"iceberg" surrounds the solute (Frank and Evans, 1945). Water also has been
considered to be an equilibrium mixture of an icelike and a close-packed
structure, and with a molecule of gas as a solute, it reacts with the icelike
structure filling one of the cavities to form a gas-hydrate and shifting the
equilibrium from the close-packed structure to the icelike structure (Namoit,
1961).
Another theory is that water is composed of clusters of highly hydrogen-
bonded molecules which are surrounded by a closely packed structure of
monomeric water. These flickering clusters form and dissolve perpetually as
a result of local energy fluctuations. Therefore, a water molecule can have a
solute molecule as a neighbor along with its four H-bonded water neighbors.
Interactions between the solute and water molecules will depress the energy
level of the tetrabonded water mo lecule. However, large numbers of water
molecules surround an unbonded molecule, and if it acquires a solute
neighbor after the latter replaces a water molecule, the energy level is raised.
Changes in the energy levels cause a shift of water molecules between various
levels in accordance with the Boltzmann distribution law, giving an increase
in the "icelikeness" and an increase in the clusters of water molecules near
the surface oí the solute molecule (Nemethy and Scheraga, 1962).
When a hydrocarbon molecule transfers from the pure liquid to the solu-
tion hydrocarbon-water, interactions are established while hydrocarbon-
hydrocarbon interactions are broken. The amount, kind, and state
(suspended, dissolved, or colloidal) of organic matter in petroleum-associated
waters is important in determining the origin and migration of petroleum,
and in prablems concerning pollution of fresh waters by petraleum-
associated waters. Prabably the most plausible theory concerning the origin
of petroleum is that it originated fram organic constituents which are
recognized as remnants or degradation products of living organisms of past
ages; these organic source material s entered fine-grained aquatic sediments
where biochemical and chemical conversions and fractionations occurred
(Erdman, 1965). As increased sedimentation took place, the resulting over-
burden pressure and compaction caused the interstitial water, which con-
178 ORGANIC CONSTITUE TS 1 SALINE WATERS

tained minute quantities of hydrocarbon, to be squeezed out of the con-


solidating sediments, and subsequently the oil was accumulated in sands and
left behind in structural traps (Kidwell and Hunt, 1958). Obviously, the
waters associated with petroleum playa very important part in the origin,
rnigration, accumulation, and subsequent production of petroleum - the
accumulation and production of petroleum being total!y dependent upon
hydraulic flow in response to geostatic and hydrostatic pressures.
Consider briefly that 99% of the oil found and produced typical!y occurs
within the pore spaces of sedimentary rocks (Hedberg, 1964). About 59o/r of
the production comes from sandstone reservoirs, 40% from carbonates, and
1% from other types of rock. Petroleum in igneous and metamorphic rocks
occurs primarily in fracture pare spaces and probably has migrated to these
rocks from its place of origino
The solubilities of petroleum hydrocarbons in water increase with
temperature and de crease as the salinity of the water increases. A tempera-
ture drop frorn 150° to 25°C reduces the solubility of petroleum in water by
a factor of 4.5-20.5. Such a mechanism can account for the accumulation
of petroleum because as upward moving subsurface waters containing
dissolved hydrocarbons decrease in temperature and pressure and meet more
saline waters, they will release dissolved hydrocarbons (Price, 1973).
Information concerning dissolved organic matter in sea water was
published as early as 1892 (Duursma, 1965). Palmitic acid, stearic acid,
acrolein, and organic nitrogen were tentatively identified. The dissolved
organic matter was found to be about 2 mg/I in the open sea, increasing to
about 15 mg/I in water taken near the coast of Greece, al! of which was
attributed to saponification of the fats of dead organic organisms. Phyto-
plankton organisms comprise most living marine organic matter, 10% of
which eventual!y becomes animal matter. The bulk of the organic particulate
matter in the sea results from dead animal matter, but the dissolved organic
constituents appear to be derived from dead phytoplankton and detritus
rather than excretions from living cells.
Decomposition of organic matter results primarily from the activities of
heterotrophic bacteria. Organic matter decomposes more rapidly in a near-
shore environment, where there is an abundance of such matter and bacteria,
than in a deep-sea environment, where both the matter and bacteria are
diluted. The dissolved organic matter can be classified into groups as follows:
(1) nitrogen-free (for example, carbohydrates); (2) nitrogen-containing (for
example, proteins); (3) lipids (for example, esters of fatty acids); and (4)
complexes comprised of mixtures of the preceding three groups (for
example, humic acids).

Nitrogen-free organic compounds

Many petroleum-associated waters contain methane; however, in Japan,


there is a type of natural-gas deposit cal!ed "suiyósei-ten'nengasu", a dry gas,
NITROGEN·FREE ORGANIC COMPOUNDS 179

which is found dissolved in subsurf'ace brines (Marsden and Kawai, 1965).


The major reservoirs in which it is found are marine or lagoonal sedimentary
basins with thick sediments and of wide areal extent. Some of the a ssociated
brines contain more than SO mgjl of iodide, which is the only commercial
source of iodine in Japan. Some of these brines also contain dissolved
ethane, propane, isobutane, butane, isopentane, and pentane.
An interesting note is that large Soviet deposits of natural gas in a solid
state totaling about 15 trillion m ' were reported to the U.S.S.R. Committee
for Inventions and Discoveries. According to U.S.S.R. investigators, mole-
cules of ground water attract molecules of natural gas and convert them to a
hydrate, which resembles silvery-grey ice, where the pressure is 250 atm and
the temperature 25°C or less. 1 m ' of the hydrate contains up to 200 m ' of
natural gas. These solid hydrate deposits are found in permafrost zones at
depths to 2,500 m. Because of the high electrical resistance, they are
discoverable by geophysical methods. The hydrate can be converted to gas
by sinking a well and reducing the pressure andjor pumping a catalyst such
as methyl alcohol into it (Anonymous, 1970).
The solubility of the hydrocarbons benzene, toluene, o-xylene, m-xylene,
p-xylene, naphthalene, biphenyl, diphenylmethane, and phenanthrene was
found to increase with increasing silver-ion concentration, indicating that a
slightly soluble 1-1 complex formed (Andrews and Keefer, 1949). Evidence
was obtained that two water-soluble complexes formed with silver and each
aromatic hydrocarbon tested. Potassium nitrate causes a reduction in the
solubility of aromatics in aqueous solutions (salting-out effect) , but silver
nitrate in creases the solubility of toluene about 73% compared to its solu-
bility in pure water. Apparently this effect with silver ions results from the
formation of rr-complexes between the benzene ring and the cation.
Benzene hydrocarbons exhibit a minimum solubility in water near lSo C
which corresponds to a zero heat of solution. The actual volume occupied by
a hydrocarbon with one benzene ring in water solution apparen~ly influences
its degree of solubility, and the larger the molecule the less soluble it is in
water (Bohon and Claussen, 1951). However, naphthalene and biphenyl,
which are larger in size and are multiring compounds, were 7 to 10 times
more soluble, indicating that some property of the benzene ring may in-
fluence the solubility. It was postulated that a positive heat of solution
resulted in the heat of cavity .Iormation, while a negative heat of solution
resulted from the formationof icelike structures around the dissolved
hydrocarbons andjor a 7r-electron complex of the aroma tic nucleus where
the 7r-electrons functioned as a base and the water as an acid. The heat of
cavitation would predominate above lSoC and would cancel the negative
heat reaction at lSoC , and below is-c the negative heat would be larger.
A study of the effect that the salts sodium fluoride, sodium chloride,
lithium chloride, ammonium chloride, sodium iodide, cesium chloride,
tetramethylammonium bromide, etc., have upon the activity coefficient of
benzene in aqueous solutions indicates that the salting-out effect varies con-
180 ORGANIC CONSTITUENTS IN SALINE WATERS

siderably among the electrolytes (McDevit and Long, 1952). A limiting law
for determining the influence of electrolytes on the activity of non polar
solutes was developed, which related the magnitud e of the salt effect to the
volume changes associated with salt and water mixing.
Molecular hydrogen was found in oilfield waters in the Lower Volga region
(Zinger, 1962). Up to 43% of the dissolved gas in these waters was hydrogen;
other gases dissolved in the water s were methane, ethane, butane, pentane,
carbon dioxide, nitrogen, helium, and argon. The pH of these water s was as
low as 3.4, and the iron content was as high as 1,100 ppm.
The solubility of methane increases with pressure and decreases with in-
creased salt conoentration at ambient temperature in NaCl-H2 O and
CaC12-H2 O systems (Duffy et al., 1961). From the experimental data, it
was estimated that 1 cubic foot of sedimentary rock with 20% porosity
buried 300 m deep and saturated with a brine containing 50,000 ppm of
NaCl could accommodate 0.3 mole of methane in solution.
A gas+liquid partition chromatographic technique was used to determine
the solubilities of C¡ -C9 paraffin and branched-chain paraffins, four
cycloparaffins, and five aromatic hydrocarbons in water (McAuliffe, 1963).
Later this study was extended to seventeen paraffins, seventeen olefins, nine
acetylenes, seven cycloparaffins, seven cycloolefins, and six aroma tic hydro-
carbons (McAuliffe, 1966). The data indicated that the solubilities of the
hydrocarbons in water increased as unsaturate bonds were added to the
molecule, with ring closure, with addition of unsaturate bonds in the ring,
and with addition of bonds which decreased the hydrocarbon molar volume.
Branching increased the solubility in water for paraffin, olefin, and acetylene
hydrocarbons but not for the cycloparaffin, cycloolefin, and aromatic hydro-
carbons. Plots of the log of the solubility in water were a linear function of
hydrocarbon molar volume for each homologous series of hydrocarbons.
A capillary-cell method was used to measure the diffusion of methane,
ethane, propane, and n-butane in water (Witherspoon and Saraf, 1964). At
25°C, the results indicated that the diffusion coefficients times 105 cm2jsec
were 1.88, 1.52, 1.21, and 0.96, respectively, for methane, ethane, propane,
and n-butane. The coefficients increased with higher temperatures.
Near the critical solution temperature and about 300 atm, the solubility
versus pressure curves for some hydrocarbon-water systems show a sharp
maximum. However, pressure has a negative effect on solubility beyond this
maximum, and a second two-phase region appears. The five binary
hydrocarbon+water systems studied were benzene, n-heptane, n-pentane,
2-methyl-pentane, and toluene (Connolly, 1967).
The accommodation of C 12----<:::36 n-alkanes in distilled water was deter-
mined as a function of hydrocarbon supply, settling time, filtration pore-
size, and mode of introduction (Peake and Hodgson, 1967). Apparently it is
possible to accommodate hydrocarbons in water at levels higher than solubil-
NITROGEN-FREE ORGANIC COMPOUNDS 181

ity levels, and such accommodation systems are stable for several days.
Preferential accommodation of alkanes in the C l6-C20 range was found at
the expense of other alkanes with lower and higher carbon numbers.
A gas chromatographic method for the determination of petral in water
was developed whereby the petral was extracted fram the water into
nitrobenzene and the extract was analyzed using a column polyethylene
glycol 1,500 on silanized Chromosorb W (Jeltes·and Veldink, 1967). The
methods were sensitive to 0.1 rng/I, and for concentrations > 0.5 mg/l the
precision was about 5% for the majar components.
Low-molecular-weight hydrocarbons in the C¡ -C4 range were detected in
sea water. Generally the concentration tended to decrease with depth
(Swinnerton and Linnenbom, 1967). Methane was the most abundant
hydrocarbon found, but smaller amounts of ethane, ethylene, propane,
propylene, n-butane, isobutane, and some buten es also were detected and
measured.
Hundreds of drill-stem samples of brine fram water-bearing subsurface
formations in the Gulf coastal area of the United States were analyzed to
determine their amounts and kinds of hydracarbons (Buckley et al., 1958).
The chief constituent of the dissolved gases usually was methane, with mea-
surable amounts of ethane, propane, and butane presento The concentration
of the dissolved hydrocarbons generally increased with depth in a given
formation and also increased basinward with regional and local variations. In
clase praximity to some oilfields, the water s were enriched in dissolved
hydrocarbons, and up to 14 standard cubic feet of dissolved gas per barrel of
water was observed in some locations.
The ratio of toluene to benzene in 27 crude oils from various sources
ranged from 2.0 to 11.3. Toluene is less soluble than benzene in distilled
water, where the ratio is about 0.3 (McAuliffe, 1966). A method of
praspecting for petroleum, utilizing information concerning the amount of
benzene dissolved in subsurface waters, was patented (Coggeshall. and
Hanson, 1956). Gas chromatographic methods proved to be good for deter-
mining the amount of benzene and other hydracarbons in the petroleum-
associated waters (Zarrella et al., 1967). Collected information indicates that
the concentration of benzene in petraleum-associated water varies with
different types of hydrocarbon accumulations, that the benzene concen-
tration decreases with increasing distance from the hydracarbon accumula-
tion, and that benzene is specific for detecting the occurrence of petroleum
hydracarbon accumulation in a given geologic horizon. A brine sample taken
fram a horizon separated by 27 m of shale fram an oil pool contained 0.02
ppm of benzene, indicating that low-permeability shale prevents movement
of hydracarbons.
Chromatographic techniques were developed for the determination of
sugars and phenols in sea waters and in sediments (Degens and Reuter,
1964). Biogeochemical differences were observed between the sugars in the
sea and in the sediments.
182 ORGANIC CONSTITUENTS IN SALINE WATERS

Wilson et al. (1970) found that ethylene, propylene, and carbon


monoxide are produced in illuminated sea water to which dissolved phyto-
plankton was added. Higher saturated gaseous hydrocarbons and methane
were not produced.
Bonoli and Witherspoon (1968) measured the diffusion coefficients of
methane, ethane, propane, n-butane, n-pentane, benzene, toluene, ethyl-
benzene, cyclopentane, methylcyclopentane, and cyclohexane in pure water
at temperatures ranging from 2° to 60° C using the capillary-cell method. The
effect of sodium chloride was studied, and the largest decrease in diffusion
coefficients was found for the paraffin hydrocarbons. They attributed the
decrease to the effects of ions in water acting as structure breakers as well as
obstacles to diffusion because of obstructions and hydrations.

Hydrocarbons containing nitrogen

Chromatographic techniques were developed for determining humic acids,


amino acids, and indoles in saline waters and in sediments (Degens and
Reuter, 1964). Arginine was found in the particulate matter in sea water and
decreased in concentration with depth. Relatively abundant concentrations
of ornithine, serine, and glycine were found in sea water.
The total concentrations of amino acids found in some petroleum-
associated waters ranged from 20 to 230 f.1g/1(Degens et al., 1964). In
general, the amino acid content increased with salinity. Adjustment of the
salinity of the brines to that of modern sea water indicated a similarity
between the amino acid spectra in the two. High concentrations of serine
and the presence of threonine and phenylalanine and glutamic and aspartic
acids were found in the petroleum-associated waters. It was postulated that
the amino acids occurred in the petroleum waters in a combined state as
nonproteinaceous acid complexes and that the solubility of these complexes
probably is a function of salinity. This postulate was based on information
which indicated that serine is thermally unstable. More recent information
indicates that serine, Iysine, threonine, glycocol, histine, isoleucine, and
leucine are fairly stable up to 180°C (Califet and Louis, 1965).
Liquid-exchange chromatography was used to determine the amounts of
amino acids in some saline waters (Siegel and Degens, 1966). The results
indicated the bulk of the amino acids dissolved in the sea are tied up in
complexes and are not in a free formo
A study of the organic solutes in sea water led to the conclusion that
coprecipitation methods are the most versatile for their isolation (Chapman
and Rae, 1967). Some of the organics that can be isolated by this method
include glucose, glutamic acid, aspartic acid, citric acid, succinic acid, glycol-
late, glycine, and Iysine. The percent of recovery of these solutes by this
method varied from 16 to 90%. The method involved the coprecipitation of
these organic solutes with iron or copper.
Most of the nitrogen in humic acid is located in the large and intermediate
FATIY ACIDS 183

size molecules, with the smallest molecules being practically nitrogen-free.


This is attributed to the fact that the link involving nitrogen is more
susceptible to oxidatíon than the rest of the molecule. Amino acids can be
released frorn humic acid by acid hydralysis, alkaline hydrolysis, and reduc-
tion with sodium amalgam (Piper and Posner, 1968).
A widely used procedure for concentrating and recovering trace organics is
the carbon-adsorption method developed by Braus et al. (1951). A modifi-
cation of this method by Robinson et al. (1967) allowed recovery of
organics using three activated carbon filters in series, with the final two
receiving acidified water.
Krause (1962) investigated the decomposition of organic matter in natural
waters and found that immediately after the death of an organism that
amino acids and keto acids appeared in the water. After 24 hours of aerobic
decomposition, a qualitative and quantitative maximum was reached by both
groups, and the amino acids present were alanine, aspartic, glutamic, glycine,
leucine, lysine, methionine, phenylalanine, serine, tyrasine, and raline; and
the keto acids present were pyruvic, oc-ketoglutaric, oxaloacetic, and
glyoxylic. After 10 days, the only acids that remained of the amino group
were glutamic, glycine, lysine, and serine; and of the keto group, pyruvic and
oc-ketoglutaric.
Litchfield and Prescott (1970) analyzed sea water, and pond water, and
spent algal media and found aspartic acid in all of the samples. Other amino
acids frequently found were serine, glycine, alanine, and arginine. Techni-
ques employed in the analysis were dansylation, extraction, and thin-layer
chroma tography.

Fattyacids

Ralston and Hoerr (1942) determined the solubilities of the normal


saturated fatty acids frorn caproic to stearic acid, whose number of carbon
atoms ranges from 6 to 18 in water, ethanol, acetone, 2-butanone, benzene,
and glacial acetic acid from 0° to about 60°C. In general, the solubilities
increased with increasing temperature.
Free fatty acids and hydroxyl ions form when soaps hydrolyze. The rate
and percentage of hydrolysis is pH dependent, generally the potassium soaps
are more hydralyzed than the corresponding sodium soaps, and free fatty
acid never separates as such frorn pure soap solutions unless reacted with an
excess of acid such as carbon dioxide (McBain et al., 1948).
Quantitative recovery of organic constituents from saline waters without
alteration is difficult. Temperature and pressure changes, bacterial actions,
adsorption, and the high inorganic/organic constituents ratio in most
petroleum-associated waters are some reasons why quantitative recovery is
difficult. Some of these factors apply also to sea waters, and Jeffery and
Hood (1958) evaluated five methods which proved effective for isolation
of portions of the soluble organic compounds in sea water. They were: (1)
184 ORGANIC CONSTITUENTS IN SALINE WATERS

dialysis or electrodialysis; (2) ion exchange; (3) solvent extraction; (4)


coprecipitation; and (5) carbon adsorption. Their results showed that the
total organic material was most efficiently removed by electrodialysis or by
coprecipitation with ferric hydroxide.
A study of the differential uptake of organic compounds by montmoril-
lonite and kaolinite revealed that montmorillonite adsorbed the compounds
more efficiently than kaolinite. The compounds studied were aspartic acid,
alanine, glucose, and sucrose (Williams, 1960). Approximately 13% of the
aspartic acid was removed from solution by montmorillonite, while kaolinite
removed only about 2%.
The following saturated, monosaturated, and diunsaturated long-chain
fatty acids were found in sea water: saturated CIO,C12,C14,C16,C18,C20,and
C;2; diunsaturated C18; and monounsaturated CI6 and CI8 (Emery and
Koerner, 1961). Also isolated were CIS' C17, and CI9 acids which might or
might not have been originally present.
A gas chromatographic method was developed for the determination of
trace amounts of the following fatty acids in water: n-valeric, isovaleric,
n-butyric, isobutyric, propionic, and acetic (Emery and Koerner, 1961). The
gas chromatograph was equipped with- a flame ionization detector and a
column of Tween 80 on Chromosorb W.
The fatty acids lauric, myristic, palmitic, stearic, hyristoleic, palmitoleic,
oleic, linoleic, and linolenic were identified in sea water using solvent extrae-
tion, esterification, and gas-liquid chromatography (Slowey et al., 1962).
Samples of deep-sea water contained less unsaturated acids and shorter-chain
acids than surface samples.
Saturated straight-chain fatty acids were found in petroleum-associated
waters from two reservoirs. The carbon numbers were C 14through C30.The
same acids were identified in a shale-core sample from a petroleum reservo ir.
The even-numbered acids predominated over the odd-numbered acids in the
amounts found in every case. The identification methods consisted of extrae-
tion by refluxing, esterification, gas chromatography, and mass spectrometry
(Cooper, 1962).
A gas chromatographic method capable of separating unesterified fatty
acids was developed (Metcalfe, 1963). Acids up to C20were identified using
a thermal conductivity detector and a column composed of phosphoric acid-
treated ethylene glycol succinate polyester on Chromosorb W.
Bordovskii (1965) studied the sources of organic matter in marine basins,
the sedimentation of organic matter in water, and the transformation of
organic matter in sediments and its early diagenesis. He also pointed out that
the organic matter in water is present in true solution, as colloidal organic
detritus, and as live organisms in suspension. Bacteria play an important part
in altering the composition of the organic material in the aqueous phase as
well as in the sediments.
Wangersky (1965) found that organic carbon was present in freshly
distilled water and that it survived triple distillation and distillation from
«

NAPHTHENIC AND HUMIC ACIDS 185

alkaline permanganate. He correlated this with algal growth in city reservoirs


and suggested that the organic carbon content of distilled water must be
considered by anyone grawing organisms in distilled water. He also found
that bubbling sea water caused organic compounds to form aggregates on the
surface and further theorized that such reactions may be related to the origin
of life.
Kabot and Ettre (1963) developed gas chramatographic methods capable
of determining free fatty acids. They analyzed different mixtures of the
normal fatty acids using both packed and Golay columns in conjunction
with a flame ionization detector. They concluded that the quantitative
analysis of free fatty acids is possible.

Naphthenic and humic acids

Davis (1968) examined the organic fractions of artesian well waters fram a
Texas oil-bearing Eocene age aquifer, using infrared and chromatographic
methods. He found that the water coproduced with oil contained 1,000
times more naphthenic acids than water located updip from the oil. He also
found a phthalic acid ester dissolved in the petroleum-associated water but
concluded that it may be common to graund waters in general.
Shaborova et al. (1961) state that "the presence in subsurface waters of
organic acids in the form of salts of various metals or in a free state indicates
a current process of leaching of organic matter from the enclosing rack. The
presence of organic acids in subsurface waters is one of the evidences for the
existence in the earth's crust of chemical pracesses of decomposition of
preserved organic matter. In turn, the organic acids are braken down into
simpler compounds by decarboxilization. It is known that decarboxilization
of organic acids is accompanied by the formation of hydrocarbons. In
nature, this pracess is a real geochemical factor. Consequently, the organic
acids and their salts that are dissolved in subsurface waters can be regarded as
one of the sources for the generation of hydrocarbons."
Using a steam distillation method, organic acids were found in concen-
trations fram 663 to 2,242 mg/l in subsurface water s taken fram a Kazhim
stratigraphic well. The average molecular weight of the acids was from 46 to
58, and the waters taken fram Devonian age sediments contained higher
concentrations of the acids than water s taken from Carboniferous age sedi-
ments.
Lochte et al. (1949) analyzed water s produced with high-pressure gas
wells and identified the following acids: acetic, propionic, isobutyric,
n-butyric, isovaleric, n-valeric, n-hexanoic, and other C6 isomers. Crude oils
were treated with ammonia solution followed by electroprecipitation of the
aqueous phase to remo ve naphthenic acids (Agaev, 1961). Further isolation
of the naphthenic acids was accomplished by heating the aqueous phase to
decompose the ammonium salts and remo ve ammonia and water.
Oden (1919) recognized fulvic acid, humus acid, and hymatomelanic acid
186 ORGANIC CONSTITUENTS IN SALINE WATERS

in soils. Later, Page and Dutoit (1930) modified the name humus acid to
humic acid. Sestini (1898) demonstrated that the humic acids are of com-
plex composition and contain ethereal and anhydride components in
addition to alkyl, hydroxyl, and ketonic groups.
Burges (1960) suggested that humic acid is a single chemical substance or
a group of similar substances, and that primarily it is nonnitrogenous.
Steelink et al. (1960) fused soil humic acids and found the following degra-
dation products: catechol, profocatechuic acid, and resorcinol. Steelink and
To11in (1962) determined the presence of two free-radical species in humic
acid using an electron paramagnetic-resonance spectrometer. They believed
that one could be a semiquinone radical and the other a quinhydrone radical.
Fulvic acids, humic acids, and hymatomelanic acids have been found in
natural water s (Wilson, 1959; Black and Christman, 1963; Packman, 1964).
The brown color, characteristic of many natural waters, is attributed to
complex organic compounds which probably are derived from water-soluble
peptizable components of soil humus.
A method that can be used to determine the organic acids in petroleum-
associated waters was published by the Natural Gasoline Association of
America (1953). The water is treated withlime water to convert the organic
acids to their calcium salts, which are titrated with a standard mineral acid.

Determination of oil in water

The following method was developed by Nalco Chemical Company (1971)


and is applicable to waters and brines where the oily matter is hydrocarbons
or hydrocarbon derivatives and a11liquid or unctuous substances that have a
boiling point above 90° C and are extractable from waters or brin es at pH 5.0
or lower using benzene, chloroform, or carbon tetrachloride.
The sample is extracted with a fluorocarbon solvent which is evaporated
off in a specially designed vessel and the residual oil measured volumetrically
in a microsyringe.

Peor-shoped to p , copocity opprox.17 mi,


offset odditionol opening so thot"popped"
somple will be retoined

1IIIIIIIIIIIIIIIIiltllllllllltllllillllll

Syringe,500 JLI, lO-Id divisions

Fig. 6.1. Microsyringe-evaporating f1ask.


DETERMINATION OF OIL IN WATER 187

Apparatus. The necessary apparatus consists of:


(1) Microsyringe - evaporating flask (see Fig. 6.1): this assembly consists
of a single-neck flask of approximately 20 ml volume which tapers opposite
and slightly offset to the neck into a microsyringe equipped with a gas-tight
Teflon-tipped plunger and calibrated to measure 0-500 ,uI.
(2) 1,000-ml pear-shaped separatory funnel.
(3) Hotplate or hot-water bath: capable of being controlled in the range oí
45°-55°C at ±2°C.
(4) 500-ml Berzelius, tall form beaker.

Reagents. The necessary reagents are:


(1) 50% hydrochloric acid solution, reagent-grade.
(2) pH paper indicating strip or pH meter.
(3) 1,1,2-trichlorotrifluoroethane (Freon TF) reagent-grade, purified,
48° C boiling point.

Sampling. Collect a composite or spot sample representative of stream to be


measured. Volume to be taken will be dependent on content of oily material
and should be in the range of 1-5 liters. Sample should be caught in glass
container.

Procedure. Extraction: adjust pH of entire sample to pH 5 or below using


hydrochloric acid added in small increments. Thoroughly mix the sample
and allow it to stand 15 minutes. Measure the volume of entire sample and
transfer to separatory funnel. Add portion of 1,1,2-trichlorotrifluoroethane
extraction fluid (see Note 1) to sample container, thoroughly rinsing any
adhering oil material. Add this and balance of fluorocarbon to separatory
funnel. Shake thoroughly for 5 minutes and let stand to separate layers.
Draw off fluorocarbon layer into suitable beaker, filtering any entrained
solids, if necessary, and warm gently to boiling point (see Note 2). Continue
boiling until volume remaining can be contained in measuring flask.
Transfer to measuring flask with fluorocarbon rinse of beaker, and
immerse the flask and contents into 500-ml beaker partially filled with water
and warmed to 65°C on hotplate or in hot-water bath. Be sure open neck of
flask is clear of upper edge of beaker (can be maintained by extension of
syringe piston). Continue until volume is reduced to that of syringe volume.
Draw fluid into syringe and increase heat slowly to remove last traces of
solvent, indicated by lack of bubbles forming in the syringe column.
Measure amount of oil material in graduated syringe using graduations
midway in syringe.
Note 1: for single extractions fluorocarbon volume should be one-tenth of
the original sample volume. In double extractions for better accuracy and
reproducibility use two volumes of fluorocarbon, each one-twentieth of the
original sample volume.
Note 2: although fluorocarbon is essentially considered nontoxic, the
- -- ----- ~- --

188 ORGANIC CONSTITUE TS IN SALINE WATERS

evaporation of the solvent should be done in a well-ventilated area or under


an exhaust hood with adequate draft to handle high-density vapors.

Calculation:
,.11 measured
'------- = mg/l oily matter
ml sample

Organic acids in oilfield brines

This method measures the bulk of the organic acid salts in oilfield waters.

Reagents and apparatus. Acetone; NaOH, 0.02N; HCl, 0.05N; acetic acid, 10
mg/I; a hotplate; and a standard pH meter.

Procedure. Pipet 25 ml of filtered brine into a 250-ml beaker; add 25 mi of


acetone and adjust the pH to precisely 6. Titrate the sample to a pH of 3.5,
and record the amount of 0.05N HCI, used in the adjustment frorn pH 6 to
3.5. Boil the solution for 5 rninutes. Cool and titrate back to pH 6 with
0.02N NaOH. Make a blank determination for NaOH and HC!. To calculate,
subtract blank from NaOH and HCI titrations.

Calculation:
(mI HCI x HCI N) - (mI NaOH x NaOH N)
x 60,030 = mg/l organic acids
m I samp Ie . id
as acetic aci s.

References

Agaev, A.A., 1961. Separation of ammonium salts of naphthenic acids from crude oil in
an electric field. Izv. Vyssh. Uchebn. Zaved., Ne]t Caz, 4 :95-98 (in Russian).
Andrews, L.J. and Keefer, R.M., 1949. Cation complexes of compounds containing
carbon+-carbon double bonds, Il. The solubility of cuprous chloride in aqueous maleic
acid solutions. J. Am. Chem. Soc., 71 :2379-2380.
Anonymous, 1970. Solid natural gas discovered. Ind. Res., 12:70.
Black, A.P. and Christman, R.F., 1963. Chemical characteristics of fulvic acids. J. Am.
Water Works Assoc., 55:897-912.
Bohon, R.L. and Claussen, W.F., 1951. The solubility of aromatic hydrocarbons in water.
J. Am. Chem, Soc., 73:1571-1578.
Bonoli, L. and Witherspoon, P.A., 1968. Diffusion of paraffin, cycloparaffin, and arornat-
ic hydrocarbons in water and some effects of salt concentration. In: Advances in Oro
ganic Geochemistry - Proe. 4th Int. Meet. Org. Ceoehem., Amsterdam. Pergamon
Press, New York, N.Y., 9:16-18.
Bordovskii, O.K., 1965. Aceumulation and transformation of organic substance in marine
sediments. Mar. Ceol., 3:3-114.
Braus, H., Middleton, F.M. and Walton, G., 1951. Organic chemical compounds in raw
and filtered surface waters. Anal. Chem., 23:1160-1164.
REFERENCES

Buckley, S.E., Hocott, C.R. and Taggart, Jr., M.S., 1958. Distribution of dissolved
hy drocarbonsin subsurface waters. In: L.G. Weeks (Editor), Habitat of Oil. American
Association of Petroleum Geologists, Tulsa, Okla., pp. 850--882.
Burges, A., 1960. The nature and distribution of humic acid. Sci. Proc. R. Dublin Soco
Ser. A, 4:53-57.
Califet, Y. and Louis, M., 1965. Contribution to the knowledge concerning the stability
of amino acids contained in sedimentary rocks. Compt. Rend., Acad. Sci. Fr.,
261 :3645-3646.
Chapman, G. and Rae, A.C., 1967. Isolation of organic solutes from sea water by
coprecipitation. Nature, 214:627-628.
Coggeshall, N.D. and Hanson, W.E., 1956. Method of geochemical prospecting. U.S.
Patent, No. 2,767,320.
Connolly , J.F., 1967. Solubility of hydrocarbons in water near the critical solution tem-
peratures. J. Chem, Eng. Data, 11 :13-16.
Cooper, J.E., 1962. Fatty acids in recent and ancient sediments and petroleum reservo ir
waters. Nature, 193:744-746.
Davis, J.B., 1968. Distribution of naphthenic acids in an oil-bear ing aquifer. Presented at
Annual Geol. Soco Am. and Assoc. Soco Meet., Mexico, D.F., Nouember, 1968.
Program, p.71.
Degens, E.T. and Reuter, J.H., 1964. Analytical techniques in the field of organic
geochemistry. In: U. Colombo and G.F. Hobson (Editors), Aduances in Organic
Geochemistry. Pergamon Press, New York, N.Y., pp.377-402.
Degens, E.T., Hunt, J.M., Reuter, J.H. and Reed, W.E., 1964. Data on the distribution of
amino acids and oxygen isotopes in petroleum brine waters of various geologic ages.
Sedimentology,3:199-225.
Duffy, J.R., Smith, N.O. and Nagy, B., 1961. Solubility of natural gases in aqueous salt
solutions, 1. Liquidus surfaces in the System CH4 -H2 O-NaCI-CaCI2 at room temper-
atures and at pressures below 1,000 psia. Geochim. Cosmochim. Acta, 24:23-31.
Duursma, E.K., 1965. The dissolved organic constituents of sea water. In: J.P. Riley and
G. Skirrow (Editors), Chemical Oceanography. Academic Press, New York, N.Y.,
1:433-475.
Eley, D.D., 1939. Solubility of gases, 1. Inert gases in water. Trans. Faraday Soc.,
35:1281-1293.
Emery, E.M. and Koerner, W.E., 1961. Gas chromatographic determination of trace
amounts of the lower fatty acids in water. Anal. Chem., 33 :146-147.
Erdman, J.G., 1965. Petroleum - its origin in the earth. In: A. Young and J.E. Galley
(Editors), Fluids in Subsurface Enuironments - Am. Assoc. Peto Geol., Mem. 4,
pp.20-52.
Frank, H.S. and Evans, M.E., 1945. Free volume and entropy in condensed systems. J.
Chem. Phys., 13:507-532.
Hedberg, H.G., 1964. Geologic áspect s of origin of petroleum. Bull. Am~ Assoc. Peto
Geol., 48 (11):1755-1803.
Jeffery, L.M. and Hood, D.W., 1958. Organic matter in sea water; an evaluation of various
methods for isolation. J. Mar. Res., 17 :247-271.
Jeltes, R. and Veldink, R., 1967. The gas chromatographic determination of petrol in
water. J. Chromatogr., 27:242-245.
Kabot, F.J. and Ettre, L.S., 1963. Gas chromatography of free fatty acids, IlI. Quantita-
tive aspects. J. Gas Chromatogr., 1:7-10.
Kidwell, A.L. and Hunt, J.M., 1958. Migration of oil in sediments. In: L.G. Weeks
(Editor), Habitat of Oil. American Association of Petroleum Geologists, Tulsa, Okla.,
pp. 790-817.
Krause, H.R., 1962. Investigation of the decomposition of organic matter in natural
waters. FAO Fish. Biol. Rep., No.34, 14 pp.
190 ORGANIC CONSTITUENTS IN SALINE WATERS

Litchfield, C.D. and Prescott, J.M., 1970. Analysis by dansylation of amino acids
dissolved in marine and freshwaters. Limnol. Oceanogr., 15 :250-256.
Lochte, H.L., Burnam, C.W. and Meyer, H.W.H., 1949. Organic acids produced by high
pressure gas wells. The Peto Eng., 21 :C34-C40.
Marsden, S.S. and Kawai, K., 1965. "Su iyósei-Tennengasu " - a special type of Japanese
natural gas deposit. Bull. Am. Assoc. Peto Geo/., 49:286-295.
McAuliffe, C., 1963. Solubility in water of C 1 -C9 hydrocarbons. Nature, 200:1092-1093.
McAuliffe, C., 1966. Solubility in water of paraffin, cycloparaffin, olefin, acetylene,
cycloolefin, and aromatic hydrocarbons. J. Phys. Chem., 70:1267-1275.
McBain, J.W., Laurent, P. and John, L.M., 1948. The hydrolysis of soap solutions, m.
Values of pH and the absence of fatty acid as free liquid or solido J. Am. Oil Chem.
Soc., 25:77-84.
McDevit, W.F. and Long, F.A., 1952. The activity coefficient of benzene in aqueous salt
solutions. J. Am. Chem. Soc., 74:1773-1777.
Metcalfe, L.D., 1963. The gas chrornatography of fatty acids and related long chain
compounds on phosphoric acid treated columns. J. Gas Chromatogr., 1 :7-11.
Nalco Chemical Cornpany, 1971. Ana/ytica/ Method [or the Determination of Oil in
Water. Unpublished method (permission obtained through H.C. Noe).
Namoit, A.Y., 1961. Solubility of nonpolar gases in water. Zh. Struht. Khim., 2 :408-417.
Natural Gasoline Association of America, 1953. Condensate Wel/ Corrosion. NGAA,
. Tulsa, Okla, 203 pp.
Nemethy, G. and Scheraga, A. A. , 1962. Structure of water and hydrophobic bonding in
proteins, 1. A model for the thermodynamic propertiesof liquid water. J. Chem.
Phys., 36 :3382-3400.
Oden, S., 1919. Humic acids, studies in their chernistry , physics, and soil science.
Kolloidchem. Beih., 11: 7 5-260.
Packman, R.F., 1964. Studies of organic color in natural water. Proc. Soco Water Treat.
Exam., 13:316-329.
Page, H.J. and Dutoit, M.M.S., 1930. Studies on the carbon and nitrogen cycles in the
soil, III. The formation of natural humic matter. J. Agric. Sci., 20:478-488.
Peake, E. and Hodgson, G.W., 1967. Alkanes in aqueous systems, 11.The accommodation
of C12-C36 n-alkanes in distilled water. J. Am. Oil Chem. Soc., 44:696-702.
Piper, T.J. and Posner, A.M., 1968. On the amino acids found in humic acid. Soi/ Sci.,
106:188-192.
Price, L.C., 1973. Solubility of petroleum in water as function of temperature and
salinity and its significance in primary migration. Bull. Am. Assoc. Peto Geo/., 57 :80l.
Ralston, A.W. and Hoerr, C.W., 1942. The solubilities of the normal saturated fatty acids.
J. Org, Chem., 7:546-555.
Robinson, L.R., O'Conner, J.T. and Engelbrecht, R.S., 1967. Organic materials in Illinois
groundwaters. J. Am. Water Works Assoc., 59:227-236.
Sestini, F., 1898. The nitrogen content associated with humic acids in soil. Landw. Verso
Sta., 51 :153-160.
Shaborova, N.T., Tunyak, A.P. and Nektarova, M.B., 1961. Study of organic acids in
underground waters. Geo/. Nefti Gaza, 5 :50-51 (in Russian).
Siegel, A. and Degens, E.T., 1966. Concentration of dissolved amino acids frorn saline
waters by ligand-exchange chromatography. Science, 151 :1098-1101.
Slowey, J.F., Jeffery, L.M. and Hood, D.W., 1962. The fatty-acid content of ocean water.
Geochim. Cosmochim. Acta, 26:607-616.
Steelink, C. and Tollin, G., 1962. Humic acid contains stable free radicals. Chem. Eng.
News, 40 :53-54.
Steelink, C., Berry, J.W., Ho, A. and Nordby, H.E., 1960. Alkaline degradation products
of soil humic acid. Sci. Proc. R. Dub/in Soc., Ser. A, 1(4):59-67.
Swinnerton, J.W. and Linnenbom, V.J., 1967. Gaseous hydrocarbons in sea water: deter-
mination. Science, 156:1119-1120.
REFERENCES 191

Wangersky, P.J., 1965. The organic chemistry of sea water. Am. Sci., 53:358-374.
Williams, P.M., 1960. Organic Acids Found in Pacific Ocean Waters. Ph.D. Dissertation,
Univ. of Calif., Los Angeles, Calif., 74 pp.
Wilson, A.L., 1959. Determination of fulvic acids in water. J. Appl. Chem., 9:501-510.
Wilson, D.F., Swinnerton, J.W. and Lamontagne, R.A., 1970. Production of carbon
monoxide and gaseous hydrocarbons in sea water: relation to dissolved organic carbono
Science,168:1577-1579.
Witherspoon, P.A. and Saraf, D.N., 1964. Diffusion of methane, ethane, propane, and
n-butane in water. Presenled at Am. Chem. Soc., Diu. Peto Chem. Meet., Chicago, I/l.,
1964, Preprints, pp.303-311.
ZarreIla, W.M., Mousseau, R.J., CoggeshaIl, N.D., Norris, M.S. and Schrayer, G.J., 1967.
Analysis and significance of hydrocarbons in subsurface br ines. Geochim. Cosmochim.
Acta, 31:1155-1166.
Zinger, A.S., 1962. Molecular hydrogen in gas dissolved in water s of oil-gas fields, Lower
Volga region. Geochemistry , 10:1015-1023.
Chapter 7. ORIGIN OF OILFIELD WATERS

The five spheres of the earth are the lithosphere (rocks), the pedosphere
(soils, till, and other surficial materials), the hydrosphere (natural waters),
the atmosphere (gases), and the biosphere (living organisms). Oilfield waters
are a part of the hydrosphere, and petroleum is a product of the biosphere.
The total volume of water in the hydrosphere is about 1,338 x 1018liters,
and about 8.4 x 1018 liters is ground water (Skinner, 1969). Most of the
water, 1,300 x 1018 liters, is in the oceans. Less than 50% of the total
ground water is in strata below 1 km. The total amount of water in sedimen-
tary rocks and associated with liquid and gaseous hydrocarbons is less than
1.3 x 1018 liters.
All of the petroleum recovered to date has been taken from oil wells
drilled into the upper 8 km of the earth's crust. The average thickness of the

Crusl

Okm
17 km
400 km

~-~""---- 1,000 km

"-<------'r----\--\-\----- 2,900 km

~----\---_+____1__tt_- 5,000 km

--+----+---+--+++-- 6,371 km

Fig. 7.1. Regions of the interior of the earth.


194 ORIGIN OF OILFIELD WATER S

earth's crust is about 17 km, ranging from 5 km under the oceans to about
35 km under the continents (Clark and Ringwood, 1964). Fig. 7.1 illustrates
the various regions of the interior of the earth, with the distance from the
surface of the crust to the center of the inner core being 6,731 km. In this
discussion we are concerned only with the crust to a distance of 0.08% of
the depth to the center of the earth.
Hydrocarbons are believed to have originated from organic material in
sedimentary material which was produced by weathering and erosion of the
earth's surface. This eroded material is carried away by water, ice, or wind
and redeposited, ultimately forming sedimentary rocks. The major sedimen-
tary, rninerals are clays, quartz, calcite, gypsum, anhydrite, dolomite, and
haiite. Most of the large bodies of sedimentar y rocks were formed in marine
environments; smaller sedimentary deposits formed in lakebeds and river
floodplains.

Definitions of some water terms

Meteoric water. White (1957) defined it as water that was recently involved
in atmosphere circulation and further that "the age of meteoric groundwater
is slight when compared with the age of the enclosing rocks and is not more
than a small part of a geologic period."

Sea water. The composition of sea water may vary somewhat, but in general
will have a composition relative to the following (in mg/I): chloride -
19,375, bromide - 67, sulfate - 2,712, potassium - 387, sodium -10,760,
magnesium - 1,294, calcium - 413, and strontium - 8.
Table 7.1 (Anonymous, 1964) gives a more comprehensive picture of the
constituents found in sea water. The analyses given in Table 7.I. are in parts
per million.

Interstitial water. Interstitial water is the water contained in the small pores
of spaces between the minute grains or units of rock. Interstitial waters are:
(1) syngenetic (formed at the same time as the enclosing rocks); or (2)
epigenetic (originated by subsequent infiltration into rocks).

Connate water. The term connate implies born, produced, or originated


together, connascent. Therefore, connate water probably should be con-
sidered to be an interstitial water of syngenetic origino White (1957) called
connate water of this definition a fossil water, i.e., water that has been out
of contact with the atmosphere for at least a large part of a geologic periodo
As White (1957) pointed out the implication that connate waters are only
those "born with" the enclosing rocks is an undesirable restriction.

Diagenetic water. Diagenetic waters are those waters that have changed
chemically and physically, both before, during, and after sediment consolida-
194 ORIGIN OF OILFIELD WATERS

earth's crust is about 17 km, ranging frorn 5 km under the oceans to about
35 km under the continents (Clark and Ringwood, 1964). Fig. 7.1 illustrates
the various regions of the interior of the earth, with the distance from the
surface of the crust to the center of the inner core being 6,731 km. In this
discussion we are concerned only with the crust to a distance of 0.08% of
the depth to the center of the earth.
Hydrocarbons are believed to have originated from organic material in
sedimentary material which was produced by weathering and erosion of the
earth's surface. This eroded material is carried away by water, ice, or wind
and redeposited, ultimately forming sedimentary rocks. The major sedimen-
tary, rninerals are clays, quartz, calcite, gypsum, anhydrite, dolornite, and
haiite. Most of the large bodies of sedimentary rocks were formed in marine
environments; smaller sedimentary deposits formed in lakebeds and river
floodplains.

Definitions of some water terms

Meteoric water. White (1957) defined it as water that was recently involved
in atmosphere circulation and further that "the age of meteoric groundwater
is slight when compared with the age of the enclosing rocks and is not more
than a small part of a geologic period."

Sea water. The composition of sea water may vary somewhat, but in general
will have a composition relative to the following (in mg/l): chloride -
19,375, bromide - 67, sulfate - 2,712, potassium - 387, sodium -10,760,
magnesium - 1,294, calcium - 413, and strontium - 8.
Table 7.1 (Anonymous, 1964) gives a more comprehensive picture of the
constituents found in sea water. The analyses given in Table 7.1. are in parts
per million.

Interstitial water. Interstitial water is the water contained in the small por es
of spaces between the minute grains or units of rock. Interstitial waters are:
(1) syngenetic (formed at the same time as the enclosing rocks); or (2)
epigenetic (originated by subsequent infiltration into rocks).

Connate water. The term connate implies born, produced, or originated


together, connascent. Therefore, connate water probably should be con-
sidered to be an interstitial water of syngenetic origino White (1957) called
connate water of this definition a fossil water, i.e., water that has been out
of contact with the atmosphere for at least a large part of a geologic periodo
As White (1957) pointed out the implication that connate waters are only
those "born with" the enclosing rocks is an undesirable restriction.

Diagenetic water. Diagenetic waters are those waters that have changed
chemically and physically, both before, during, and after sediment consolida-
SEDIMENTARY ROCKS 195

TABLE 7.1

Average composition of sea water

Element Amount Element Amount Element Amount


(ppm) (ppm) (ppm)

Chlorine 18,980 Zinc 0.01 Tungsten 1x 10-4


Sodium 10,560 Molybdenium 0.01 Germanium 1x 10-4
Magnesium 1,270 Selenium 0.004 Xenon 1x 10-4
Sulfur 880 Copper 0.003 Chromium 5 x 10-5
Calcium 400 Arsenic 0.003 Beryllium 5 x 10-5
Potassium 380 Tin 0.003 Scandium 4 x 10-5
Bromine 65 Lead 0.003 Mercury 3x 10-5
Carbon 28 Uranium 0.003 Niobium 1x 10-5
Oxygen 8 Vanadium 0.002 Thallium 1x 10-5
Strontium 8 Manganese 0.002 Helium 5 x 10-6
Titanium 0.001 Gold 4 x 10-6
Boron 4.8 Thorium 0.0007 Praseodymium 2 x 10-7
Silicon 3.0 Cobalt 0.0005 Gadolinium 2 x 10-7
Fluorine 1.3 Nickel 0.0005 Dysprosium 2x 10-7
Nitrogen 0.8 Gallium 0.0005 Erbium 2x 10-7
Argon 0.6 Cesium 0.0005 Ytterbium 2 x 10-7
Lithium 0.2 Antimony 0.0005 Samarium 2 x 10-7
Cerium 0.0004 Holmium 8 x 10-8
Rubidium 0.12 Yttrium 0.0003 Europium 4 x 10-8
Phosphorus 0.07 Neon 0.0003 Thulium 4 x 10-8
Iodine 0.05 Krypton 0.0003 Lutetium 4 x 10-8
Barium 0.03 Lanthanum 0.0003 Radium 3 x 10-11
Indiurn 0.02 Silver 0.0003 Protactinium 2 x 10-12
Aluminum. 0.01 Bismuth 0.0002 Radon 9 x 10-15
Iron 0.01 Cadmium 0,0001

tion. Some of the reactions that occur in or to diagenetic waters include


bacterial, ion exchange, replacement (dolomitization), infiltration by
permeation, and membrane filtration.

Formation water. Formation water as here defined is water that naturally


occurs in the rocks and is present in them immediately before drilling (Case,
1955).

Juvenile water. Water that is in primary magma or derived from primary


magma is juvenile water (White, 1957).

Sedimentary rocks

At least a portion of the water found in petroleum reservoirs consisting of


sedimentary rocks was deposited with the sediments before they were trans-
formed into rock. As the sediment compacted to form rock, the composition
196 ORIGIN OF OILFIELD WATERS

TABLE 7.II

Average composition of igneous and some types of sedimentary rocks (ppm)

Element Igneous rocks Sedimentary rocks Evaporites


(halite)
resista tes hydrolyza tes precipitates

Si 277 ,200 367,500 272,800 24,200 90


Al 81,300 25,300 81,900 4,300 20
Fe 50,000 9,900 47,300 4,000 11
Ca 36,300 39,500 22,300 304,500 930
Na 28,300 3,300 9,700 370 325,000
K 25,900 11,000 27,000 2,700 800
Mg 20,900 7,100 14,800 47,700 460
Ti 4,400 960 4,300 0.8
P 1,180 350 740 175
Mn 1,000 trace 620 385 1
F 600-900 510 250 20
S 520 2,800 2,600 1,100 770
C 320 13,800 15,300 113,500 70
CI 314 trace 200 586,000
Rb 310 273 300 O
Sr 300 < 26 170 425-765
Ba 250 170 460 120 2
Cr 200 68-200 410-680 2
Zn 132 < 20 200-1,000 < 50 1
Ni 80 2-8 24 O O
Cu 70 192 20.2 3
Li 65 17 46 < 26
N 46
Sn 40 40
Co 23 O 8 O
Pb 16 20 20 5-10 2
Th 11.5 (? ) 6.1 10.1 1.1 < 0.2
Cs 7 12
Be 6 (?) <4 O
As 5 - 5
U 4 1.2 1.2 1.3 0.01
B 3 9-31 310 3 <2
Br 1.62 <0.2 60
1 0.3 0.07-0.55 <2
Cd 0.15 O 0.3
Se 0.09 0.6 < 0.1 <0.5
Hg 0.08-0.5 0.1 0.3 0.03
Ra 1.3 x 10-6 0.7 x 10-6 1.08 x 10~ 0.42 x 10~

of the interstitial water changed because of reactions with the rock. A


simplistic view of sedimentary rocks and their relation to oilfield waters
should include consideration of weathering; erosion; transportation
mechanisms; sorting of weathered products; depositional environments of
clastics, carbonates, evaporites, organic matter, and silica; sediment com-
pactions; sediment diagenesis; and petroleum and natural gas.
SEDIMENTARY ROCKS 197

The volume of the earth is about 1,100 billion krn' and the volume of the
oceans is about 1.3.billion krn ", however, the oceans with an area of 360
million km2 cover 70% of the surface of the earth. The average composition
of some of the igneous and sedimentary rocks of the earth's crust is shown in
Table 7.II, which was taken from Clarke and Washington (1924) and
Rankama and Sahama (1950). The resistate rocks referred to in Table 7.II
are composed of residues not chemically decomposed in the weathering of
the parent rocks. Hydrolyzate racks are the insoluble products formed by
chemical reactions during weathering of parent rocks. Precipitate racks are
those formed by chemical precipitation of mineral s from aqueous solution.
Evaporite rocks are marine evaporites which were produced when the water
in which they were dissolved was evaporated.
Sedimentary rocks comprise about 5% of the lithosphere, while the
igneous racks form 95%. The three main types of sedimentary rocks are
shale, sandstone, and iimestone, and their relative abundance determined
fram geochemical data ranges from 70 to 83% shale, fram 8 to 16% sand-
stone; and from 5 to 14% limestone (Pettijohn, 1957). Levorsen (1966)
noted that oil and gas are found in reservo ir rocks consisting primarily of
sandstones, limestones, and dolomites.

Weathering

Weathering is a most important factor in praducing the source material for


the creation of sedimentary rocks. Processes that cause weathering are
chemical, physical, and biological (Ross, 1943).
The weathering of rack by physical methods includes temperature changes
brought about by climate changes. Examples are the breaking of rack by
thermal expansion (heat), the breaking of rack by the expansion of freezing
water in the pares or cracks, or the mechanical breaking of rack as a glacier
moves over it. Breaking the rack causes the surface area to become larger
without significantly changing the chemical composition.
Biological weathering includes the cracking of rack as a result of plant
roots and the action of acids derived from plants, animal s, and bacteria. The
biotic factor includes bacteria, algae, protista, protozoa, higher animals and
plants, during both life and subsequent necrotic decomposition which
furnish H+ ions, colloids, complexing agents, and dispersants.
Chemical weathering involves the action of water upon the parent rack
and upon the products of physical and biological weathering. In chemical
weathering the composition of the source rack is changed by solution,
hydralysis, oxidation, and reduction reactions. The H+ ion when concen-
trated in aqueous solution is a very important energy factor because it will
cause rapid chemical reactions with parent rocks. The redox potential in-
fluences the rate of removal of elements, such as iron and manganese from
the parent rack, and if it is a reducing potential, these elements are more
likely to remain in solution after solubilization.
198 ORIGIN' OF OILFIELD WATERS

Erosion

Erosion is the opposite of deposition (the processes are reversible), but


erosion must occur before deposition can proceed. The products of weath-
ering are eroded and transported to a new location by the action of water
and wind. The water serves to transport the majority of these products, and
it can transport them by dissolution, suspension, or pushing of larger
particles.

Transportation mechanisms

Both wind and water can transport the products of weathering, however,
this discussion will consider only water. The transport mechanisms con-
sidered are chernistry, physics, and hydraulics.
Perhaps the primary solvents of weathered products are carbonated water,
organic acids, and sulfate solutions. Elements that dissolve readily in car-
bonated waters are lithium, sodium, potassium, magnesium, calcium, stron-
tium, iron, manganese, phosphorus, and others. The organic acids will
dissolve iron and manganese, while sulfate solutions will dissolve copper,
iron, and manganese compounds.
The chemistry of the water is a prime factor in the dissolution of the rock;
if the pH is acidic, the transition group metals are more likely to dissolve,
while if it is basic, elements such as silica are more likely to dissolve. Salts of
the alkali and alkaline earth metal s will dissolve if the pH is either acidic or
basic; however, if the pH is above 10, some of the alkaline earths such as
magnesium will precipitate. The pH of the water is influenced by the dissolu-
tion of carbon dioxide. For example, as carbon dioxide dissolves in water,
the pH will change. The pH of pure water in equilibrium with carbon dioxide
can be calculated and is 5.65 (Hem, 1970). The pH is calculated using the
mass-law equations in which the activity of water is unity in dilute solution,
and kw = constant equal to the product of the activities of H+ and OH-.
Introduction of another phase such as calcite into the water carbon
dioxide system will change the pH. Garrels and Christ (1965) calculated that
such a system in equilibrium with the atmosphere will have a pH of 8.4.
Additional ions such as those found in ocean water will produce other pH
values. For example, if the system is ocean water in equilibrium with carbon
dioxide, the pH at each equilibrium step is approximately:

H20 + CO2 H2C03 l(pH 5)


H2C03 HC03 - + H+ (pH 6.3)
HC03 + H+ 2H+ + C03-2 (pH 10.3)

Some chemicals when dissolved in water act as buffers, where a buffer is


defined as something that .produces an effect which inhibits a large pH
change when an acid or a base is added to the water. Therefore, as a water
200 ORIGIN OF OILFIELD WATERS

Sorting o] weathered products

Products of weathering are transported by water by rolling along the


streambed, by suspension of the smaller particles, and by solution of soluble
components. As the larger rocks and pebbles rol! along the streambed, they
are abraded by bumping against each other and by abrasive action with the
rocks in the streambed. The size of the clastics, which are detritus trans-
ported mechanically to the point of sedimentation and portions solubilized
by water before sedimentation, decreases in the downcurrent direction
(Pettijohn, 1957), and this change in grain size is primarily a sorting effect.
The effect is noted in both fluvial and marine deposits.

Mobile belt

Geosyncllnol trough Borderlond

Forelond
(stoblel Shel f oreo

Fig. 7.3. Idealized depositional basin. (After Moore, 1969.)

Knowledge of the sorting of the clastics is used in reconstructing the


ancient environment (Visher, 1965).This knowledge can be applied to ex-
ploration for petroleum and other valuable minerals. A simplistic deposi-
tional basin is shown in Fig. 7.3; the deposited clastics will be found on the
borderland side of the basin and not on the foreland side.

Depositional enuironments of clastics

Depositional environments of the clastics include eolian, fluvial, regressive


marine, transgressive marine, deltaic, bathyal-abyssal, and lacustrine. Eolian
deposits are sands that are drifted and arranged by currents of air or wind.
Fluvial deposits are those related to streams, rivers, and ponds. Water in
these environments usually contains less than 10,000 mg/l of dissolved
solids.
Regressive marine deposits are land-derived sediments that are transported
seaward and settle in the ocean. The salinity of the water transporting these
sediments will vary, it is fresher at its source and becomes more brackish as it
nears the sea. The dissolved solids in contemporary sea water are about
35,000 mg/I, while some estuary waters contain about 20,000 mg/l of di s-
solved solids.
Transgressive marine deposits usually are small in volume compared to
fluvial and regressive-marine deposits. Such deposits are formed mainly by
SEDlMENTARY ROCKS 201

erosion and redeposition of sediment deposits. The salinity of the sea may
change somewhat during transgression but probably not mucho
Deltaic deposits result from a combination of environmental factors and
are related to both fluvial and regressive mar ine processes. During flood
times the rivers transport tremendous volumes.of material, both clastic and
organic, into delta areas. Deltaic deposition is a very important factor in the
formation of petroleum because of the tremendous amount of organic ma-
terial deposited.
Bathyal-abyssal deposits are formed in deep-water areas in the sea, and
turbidity currents are responsible for most of the clastic deposition (Ernery,
1960). Lacustrine deposits are those that are formed in lakes. If the lake is a
fresh-water lake, the dissolved solids may be less than 1,000 mg/l: in a
salt-water lake, the dissolved solids may be greater than 35,000 mg/l.
Consider a simplistic sedimentation area where the borderland area is the
prime source of sediments. The coarse- to fine-grained clastics which are
weathering products of the high-mountains borderland are deposited near
their source. The clastics are detritus transported mechanically to the point
of sedimentation and are not solubilized by the water before deposition.
Primarily they are the sandstones and shales (clays). They will not be found
on the foreland side of a depositional basin. Clay deposition can be detrital
or authigenic; illite often is detrital. There are at least two dozen clay
minerals, many of which occur in very minute grains and most of which
cannot be resolved 'by high-power petrographic microscopes. The electro n
microscope, X-ray diffraction, and differential thermal analysis are used to
determine the type of clay.
The clays are very important in relation to petroleum and gas because
they are the major component in the shales from which petroleum and gas
are generated. The clays also possess base exchange properties which will
react with constituents in water and petroleum. The detrital clays settle in
low-energy waters and they settle more rapidly from a saline water than from
a fresh water.

Depositional enuironments of the carbonates

Limestones and dolomites are the dominant carbonate reservo ir rocks,


while the sandstones are the dominant clastic reservoir rocks (Ham, 1962).
The carbonates were precipitated at the place where the rocks first formed,
while clastics were primarily transported grains.
Plumley et al. (1962) classified the carbonates according to an energy
index of the water from which they precipitated. They divided them into
five types. Type I is deposited in quiet water; it consists of calcite, 15-50%
clay, and < 5% detrital quartz. Type Ir is deposited in intermittently
agitated water and consists of calcite, < 25% clay, and < 50% detrital quartz.
Type III is deposited in slightly agitated water and it consists of calcite with
up to 50% detrital quartz. Type IV is deposited in moderately agitated water
202 ORIGIN OF OILFIELD WATERS

and consists of calcite with up to 50% detrital quartz. Types III and IV are
similar in percentage of minerals; however, they are further differentiated
according to grain size, sorting, roundness, and fossils, as are the other types.
Type Vis deposited in strongly agitated (high-energy) water and it consists of
calcite, < 5% clay, and < 25% detrital quartz.
The ca1cium in the carbonates is released during rock weathering and goes
into solution as bicarbonate. The solubility of calcium carbonate in water is
also dependent upon the amount of carbon dioxide in solution. If the
amount of dissolve carbon dioxide decreases, calcium carbonate is precipi-
tated; therefore, the amount of calcium carbonate precipitation increases in
warm water because the amount of carbon dioxide in solution is less than in
cold water. Considerable amounts of carbonate precipitation occur in warm
environments (Illing, 1954). Aquatic plants also absorb carbon dioxide and
cause carbonate precipitation. The deposited carbonates can be pure or
mixed with sand, clay, iron, manganese, and organic matter.
Modern carbonate deposition occurs as deep-water oozes, and reefs, and
on shallow shelves. The deep-water oozes form along the flanks of ocean
basins at depths of less than 6,000 m. They do not form at depths greater
than 6,000 m because the ca1cium carbonate solubility increases with the
increased pressure. On the flanks of the basins, terrigenous material mixes
with the calcium carbonate. Often the mixture is 65-89% percent ca1cium
carbonate with silt making up the remainder (Gevirtz and Friedman, 1966).
Reef carbonates develop in open oceans on shallow platforms forming
atolls, as isolated patches on the ocean shelves, or along the margins of shelf
areas as fringing reefs. Fringing reefs generally occur in tropical regions on
the western side of the ocean basin. Reefs form as a result of living organisms
which form 'the framework of the sediment (Ginsburg and Lowenstam,
1958).
Present-day shelf carbonates are developing in Florida Bay, on the Bahama
Banks, on the Australian shelf, and on shelves off British Honduras, Yucatan,
and India. The precipitation often occurs as shallow carbonate mud banks.
The sediment in the shallow shelf areas often consists of about 10% skeletal
material mixed with oolites, mud aggregates, grapestone, aragonite needles,
ca1careous algae, etc. The average rate of carbonate accumulation on the
Bahama Banks is 50 mg cm"? yr"" (Broecker and Takahashi, 1966). The
salinity of the water ranges from 36,000 mg/I of dissolved solids to 46,500
mg/l. The more saline waters occur in lagoonal areas. The pH ranges from
about 8.0 to 8.2 and is lowest at the end of the day because of CO2
extraction from the water by marine plants.
In the simplistic depositional basin shown in Fig. 7.3 (Moore, 1969),
limestones and reef limestones will be deposited on the foreland side, formed
from water soluble constituents that precipitated from the saline solutions.
SEDIME TARY ROCKS 203

Depositional environments of evaporites

Calcium carbonate precipitates from sea water after it begins to evaporate


(Usiglio, 1849). Removal of calcium ions from solution increases the Mg/Ca
ratio in the residual brine, and the precipitated calcium carbonate reacts with
the magnesium enriched brine to form dolomite (Deffeyes et al., 1964). The
common arder of evaporite deposition in a basin cut off from the open sea is
limestone > dolomite > gypsum > halite > potash. Evaporite deposition can
be stopped in a basin by a change in climatic regimen or tectonism. If it were
changed and new water were allowed to enter, the already deposited salts
probably would be effectively protected by the superadjacent high-density
brin e, and the lighter waters either meteoric or sea water would float on top,
developing euxinic conditions similar to a situation found in the Black Sea.
Toxic conditions in euxinic areas tend to preserve deposited organic
matter. The preserved organic matter later can be transformed to hydrocar-
bons and petroleum. Hypersaline lagoons often are very prolific in the
production of organic matter (Phleger and Ewing, 1962). Algal pads and an
abundance of organic organisms are characteristic of a high-pH, high-salinity
environment (Carpelan, 1957) such as exists before euxinic conditions
develop.
The deposition of evaporites occurs when water evaporates under re-
stricted environmental conditions. The restrictions usually are caused by
tectonism such as an uplift, regression of the strand line leaving a relict sea,
or biological building of a reef.
Sloss (1953) outlined five environments from which evaporites deposit;
they are normal marine, euxinic, penesaline, saline, and supersaline. Primary
carbonates and dolomitized carbonates precipitate from the normal marine
environment where the water contains about 35,000 mg/l of dissolved solids.
Limestones rich in organic matter and black shales often are related to
euxinic environments.
In the penesaline environment carbonates and primary dolomite formo
The salinity of the water is toxic to normal marine life but not sufficiently
saline so that halite precipitates. The dissolved solids in the water range from
about 250,000 mgjl to 350,0.00 mgjl.
From a saline environment carbonates, sulfates, and halites will precipi-
tate. From a supersaline environment potassium compounds will precipitate,
and the amount of dissolved solids in the solution will be about 500,000
mgjl.
A simplistic model of an evaporite basin is a closed basin initially filled
with sea water and evaporated to dryness. In the isochemical system a layer
of calcium carbonate is deposited over the entire basin floar as evaporation
proceeds. Next, gypsum is deposited and when the water is reduced to about
10% of the original volume, halite precipitates. Halite will deposit only in the
deeper parts of the basin and if the solution goes to dryness the more soluble
salts will deposit in any remaining depressions and on top of the halite.
204 ORIGIN OF OILFIELD WATERS

Deffeyes et al. (1964) proposed that dolomitization of limestone results


from the evaporation of sea water and precipitation of gypsum causing the
ratio of magnesium to calcium in the water to increase. The concentrated
water flows downward, because it is more dense, into the underlying sedi-
ments where it reacts with limestone to form dolomite.
Modern evaporite deposits are thin and cover relatively small areas of the
earth; however, the ancient environments indicate that these depositions
were widespread in the United States (Krumbein, 1951) and in the world
(Lotze, 1938). The majority of the major evaporite bodies are of marine
origin, and range in age frorn Cambrian to Tertiary. They form in arid marine
climates where water lost by evaporation equals or exceeds that supplied by
rainfall, rivers, or the open sea. They also form in deep-water environment
(Brongersma-Sanders, 1971).
The data in Table 7.III illustrate how the concentrations of some of the
dissolved constituents change as sea water is evaporated to dryness (Collins,
1969a). As sea water is evaporated to dryness and the chloride concentration
approaches 178,000 mg/I, calcium sulfate precipitates. Halite precipitates
when the chloride concentration approaches 275,000 mg/I, magnesium
sulfate precipitates during the next stage as the chloride concentration
approaches 277 ,000mg/l. The concentrations of the ions in solution at these
various stages approximate those shown in Table 7.IIl. The data in Table
7.IlI indicate that, as sea water evaporates, the concentrations of lithium,
magnesium, boron, chloride, bromide and iodide in the residual liquor in-
crease, and that the concentrations of sodium, potassium, rubidium, calcium,
and strontium decrease. In most depositional areas, the brines never reached
the concentration necessary for the deposition of potassium and magnesium

TABLE 7.III

Concentration changes dur ing evaporation of sea water and brine

Element Sea water CaS04 i NaCI i MgS04i KCI i MgCl2 i

Lithium 0.2 2 11 12 27 34
Sodium 11,000 98,000 140,000 70,000 13,000 12,000
Potassium 350 3,600 23,000 37,000 26,000 1,200
Rubidium 0.1 1 6 8 14 10
Magnesium 1,300 13,000 74,000 80,000 130,000 153,000
CaJcium
Strontium
400 1,700 100 10
° °
Boron
7
5
60
40
10
300
1
310 750 ° °
850
Chloride 19,000 178,000 275,000 277,000 360,000 425,000
Bromide 65 600 4,000 4,300 8,600 10,000
Iodide 0.05 2 5 7 8 8

Total 295,0,00 517,000 469,000 538,000 602,000


SEDIMENTARY ROCKS 205

salts; or if they did, these salts were later removed by leaching so that their
occurrence is relatively rare.
Holser (1963) analyzed some brine inclusions in halite from Permian age
evaporites. He found that the Br/Cl and Mg/Cl ratios in many of the brine
inclusions are similar to those found in the late stages of halite deposition.
He concluded that some of the inclusions were connate bitterns with few
diagenetic changes, and that the Br/Mg ratio of sea water has remained
relatively constant since Permian time. Some diagenetic changes were evident
in a few of the inclusions in which a large ratio of Ca/Cl and a low ratio of
804 /Cl compared to sea water were found.
8ediments commonly associated with evaporites are red beds, quartzose
sandstones, subgraywacke sandstones, carbonate rocks, and marine shales
(Krumbein, 1951). Normal marine evaporite successions are found in inter-
cratonic basins such as the Michigan and Williston Basins. Euxinic black
shales sometimes are associated with evaporites. Low redox potentials have
been found in modern evaporite (Morris and Dickey, 1957; Quaide, 1958).
Examples of modern depositional evaporites are the Karaboghaz Gulf on the
eastern side of the Caspian Sea, the Great Bitter Lake of 8uez, the Rann of
Cutch in northwest India (Grabau, 1920), and the Persian Gulf sabkhas
(Evans et al., 1963; Butler, 1969).

Deposition of organic matter

The organic matter can be biogenic (produced by living systems) or


abiogenic (not produced by living systerns). The source of biogenic matter
can be both terrestrial and marine; for example, considerable plant and
animal debris is collected from the land by streams and rivers and carried to
the sea, while in the sea large quantities of plant and animal matter live and
die.
The organic matter that is deposited with sediment usually decomposes if
the conditions are right; however, if the environment is reducing some of it
may be preserved. The preserved organic matter is transformed into other
organic compounds (Kvenvolden, 1964). During sediment diagenesis the
organic matter is transforrned to insoluble organic matter (kerogen) and
soluble petroleum hydrocarbon (Hunt and Jamieson, 1958). Chemical,
bacterial, and catalytic reactions are involved in these conversions. Ternpera-
ture and pressure affect the reaction rates. Some of the chemical reactions
are as follows:

(1) Oxidation: C2 Hm + O2 ---+ nC02 + Y2mH2 O


(2) Reduction: R'OH + H2 ---+ R'H + H20
(3) Elimination: R'COOH ---+ R'H + CO2
(4) Polymerization: (small units) -+ big molecule
(5) Cracking: C-C-C-C-C-C 1:0 C-C-C- + C-C
206 ORlGIN OF OlLFIELD WATERS

The organic matter produced by photosynthesis in the oceans is estimated


to be sufficient to produce 11 million metric tons of hydrocarbon precursors
annually (Riley, 1944). A very small amount of this organic material pre-
served in sedimentary racks each year thraugh geologic time would supply
all of the known oil and gas fields plus many undiscovered giant fields.
Shales consisting of organic material, siltstones, claystones, and limy mud
mixtures are found in the trough of a basin (Fig.7.3.). A simplistic idealized
view of the trough area is that organic matter deposited in the traugh is
preserved in the stagnant low-Eh environment. If rapid subsidence occurs,
the organic material later is transformed into hydrocarbons which move up
and out of the traugh into stratigraphic traps on the foreland side or struc-
tural traps on the borderland side of the basin.
The time interval between deposition of the organic material and con-
version to petroleum is millions of years, during which time the trough or
ocean basin is filled with sediment and buried, and the sediment is corn-
pacted to rack. Some of the water in which the sediments deposited will
remain in the rocks as interstitial water.

Deposition o] silica

Most of the silica in sediments is of the detrital variety but some is


authigenic. Silica is dissolved by waters with high pH potentials and precipí-
tated from water with a low pH. The precipitated silica often acts as a
cement.

Sediment compaction

Sediments compact or consolidate in response to an imposed load, and in


the natural environment, the load is the weight of overlying sediments
(Weller, 1959). Compaction of a sediment results in a reduction of the
interstitial volume concurrent with expulsion of interstitial water and defor-
mation of the sediment skeleton (the salid granular framework exclusive of
bound interstitial water). The grains and the interstitial water are almost
incompressible, and the rate of expulsion of interstitial water is about
identical with the rate of compaction (Taylor, 1956).
Terzaghi and Peck (1968) studied the expulsion of pare water from un-
consolidated clays, and determined that the rate at which clay compaction
occurs is dependent upon the clay permeability, its volume compressibility,
and the square of the thickness of the bed which is compacted. Shales
decrease in porosity when compacted; for example, their porasity can
decrease about 80% as they compact during burial.
White (1957) noted that very large quantities of water are removed from
sediments during their compaction. For example, water-saturated shale
decreasing in porosity fram 20 to 10% loses 1011 liters of water per km ' of
sedimento Such sediment could yield per krn' a water supply of 20 liters per
SEDIMENTARY ROCKS 207

minute for 10,000 years. Compaction water is further considered in a sub-


sequent chapter concerned with the accumulation of petroleum.
The resistance to flow of bound interstitial water is greater than the
resistance to deformation of the sediment framework during the first stages
of compaction. As pressure is increased and sediment compacts, the se di-
ment framework increases resistance to deformation, which al so governs the
deformation rate of the bound-water films and the outflow of the interstitial
water. Rosenqvist (1962) found that bound water has a higher viscosity than
unbound insterstitial water. Permeability decreases with compaction, and
this together with the increased water viscosity leads to additional resistance
to expulsion of bound interstitial water during subsequent compaction.
Porosity decreases during compaction, and at infinite depth it would
become infinitely small. Porosity and permeability are two important factors
in determining the amount of petroleum and/or water that can be recovered
from a given reservoir 01' aquifer (Pollard and Reichertz, 1952; Caraway and
Gates, 1959). The porosity of an aquifer indicates how much fluid the
aquifer can hold, and the permeability indicates how fluid can move through
the aquifer. If the porosity is high but the permeability is low, the reservo ir
may contain large amounts of oil, gas, or water, but they cannot be
recovered unless special techniques are used to increase the permeability.

Sediment diagenesis

Mineralogical and chemical changes occur in the sediments as they com-


pact. The mineralogic composition of recent marine sediments and ancient
marine sedimentary rocks are different, as is the composition of the intersti-
tial water in the recent sediments compared with the waters in ancient
stratigraphic units. Chemical reactions occur between the sediments and
their interstitial water (Chave, 1960). It has been shown that chemical
changes can be measured in recent sediments, e.g., below the sediment-
water interface, changes in pH and Eh result from degradation of sulfate ions
by bacteria (Emery and Rittenberg, 1952).
Numerous chemical inhomogeneities occur within a single core sample of
recent sediments (Degens and , Chilingar, 1967). The magnesium concen-
tration in the interstitial wáter decreases slightly with depth, while the con-
centrations of calcium and potassium increase (Siever et al., 1965). The
interstitial waters from continental shelf sediments have higher chloride con-
centrations and higher ratios of Ca/Cl, K/Cl, and Rb/Cl than the overlying sea
water; the ratios of Li/Cl and Mg/Cl are about the same as in the overlying
sea water except that the Li/Cl ratios are higher in the innershelf samples
than in the outershelf samples. The Sr/Cl ratio is higher in the overlying sea
water, while the pH and Eh values are lower in the sediments than in the
overlying sea water (Friedman et al., 1968).
A detailed study indicates that virtually no environment exists on or near
the earth's surface where the plI/Eh conditions are incompatible with
208 ORIGI OF OILFLELD WATERS

organic life (Baas Becking et al., 1960). Because CO2 is the main byproduct
of organic oxidation and the building material of plant and much bacterial
life, it plays a dominant roleo Carbon dioxide dissolves in water, producing
the bicarbonate ion and a free hydrogen ion. The concentration of the
hydrogen ion is 10-7 equiv./l (pH 7) at 20°C in pure water, but when
saturated with CO2 it rises to lO-s (pH 5). This reaction moves to the right
with increasing temperature in a closed system. In the presence of organic
constituents, the equilibrium is modified, and the pH range can extend from
2 to 12.
The ionic potentials of the constituents involved in diagenesis are impor-
tant (Cloke, 1966). Those that stay in true ionic solution up to rather high
pH levels are a", K+, Mg+2, Fe+2, Mn+2, Ca+2, Sr+2, Ba+2, etc.; they are
the soluble cations, and their ionic potentials range from O to 3, where the
ionic potential is the ratio between the ionic charge and the ionic radius.
Constituents that are precipitated by hydrolysis are those with ionic poten-
tials from 3 to 12 and include such ions as Al+ 3, Fe" 3, Si+4, and Mn+4.
Constituents that form soluble complex ions and usually go into true ionic
solution include B+3, C+4, N+s, p+s, S+6, and Mn+7; their ionic potentials
are over 12. In general, the hydroxides of the soluble cations possess ionic
bonds; therefore, they are soluble, the hydrolysates or those precipitated by
hydrolysis form hydroxyl bond s, and the soluble complex ions have hydro-
gen bonds.
Organisms that consume oxygen cause a lowering of the redox potential,
and in buried sediments it is the aerobic bacteria that attract the organic
constituents and remo ve the free oxygen from the interstitial water. Sedi-
ments laid down in a shoreline environment often differ in degree of oxida-
tion from those laid down in a deep-sea environment (Pirson, 1968). For
example, the Eh of the shoreline sediments may range from -50 to O m V
while the Eh of deep-sea sediments may range from -150 to -100 mV. The
aerobic bacteria die when the free oxygen is totally consumed, and the
anaerobic bacteria attack the sulfate ion which is the second most important
anion in the sea water. During this attack, the sulfate is reduced to sulfite
and then to sulfide. Also the Eh drops to -600 mV (Fig.7.2). Sulfide is
liberated and CaC03 precipitates as the pH rises above 8.5 (Dapples, 1959).
Sulfur has two stable isotopes, 32 S and 34 S, with a mass differential of 670.
The isotopes are fractionated during the change of S04 -2 to S-2, and S-2
is enriched in the more energetic 32S isotope. The average ratio of 32S/34S
in normal sea water sulfate is about 21.76 (Ault, 1959). The sulfate isotopes
are useful in interpreting ancient diagenetic stages.
Reactions occur during sediment diagenesis and affect the composition of
the interstitial water. Calcite is precipitated if the pH is high, or it is dis-
solved if the pH is low. Dolomitization occurs as follows:

2CaC03 + MgCl2 ~ CaMg(C03 h + CaCl2


SEDIMENTARY ROCKS 209

Other exchange reactions occur, whereby montmorillonite transforms to


illite (Burst, '1969). Ions are adsorbed by negatively charged clays from the
sea water (Krauskopf', 1956). Dissolved salts hydrolyze; e.g., olivine hydrol-
yzes to serpentine as follows:

Hydration and dehydration reactions occur, such as the gypsum-anhydrite


relation:

As depth of burial increases, many mineralogical changes take place in the


sediments. In the Gulf Coast Tertiary, the clay mineral montmorillonite
gradually disappears at depths between 2,500 and 3,000 m (Burst, 1969). It is
replaced by rnixed layer and illitic clay minerals. This change involves
chemical alteration and al so the release of water of crystallization. This
change appears to be temperature dependent with the reactions starting at
about 100°C.
Other mineral changes occur in the sandstones such as the deposition of
secondary silica overgrowths on the quartz grains causing resultant loss in
porosity. Available data differ on the loss of porosity with burial depth
(Atwater and Miller, 1965; Philipp et al., 1963). Much of the silica may
come from outside the porous sand, such as from shales whose pare water is
traversing the sand as a result of compaction. Authigenic clay minerals such
as kaolinite also form in the pares.
Interstitial water in deeply buried sediments sometimes is at a pressure
clase to the weight of the overburden. This pressure may be sufficient to
burst the rack and allow the water to move out through the fissures which it
forms. These fissures are principally vertical and extend upward into zones
of lower temperature .. The water forming and subsequently filling the fis-
sures usually is a mixture of salty interstitial sedimentary pare water and
water from the clay minerals released by their recrystallization. If it is hot,
and saturated with silica and other minerals, it could force its way upward
and as it cools deposit quartz, feldspar, calcite, and other minerals. Possibly
many hydrothermal ore veins were formed by interstitial sedimentary waters
rather than by "juvenile" waters.
These hydrothermal veins contain metallic minerals composed of corn-
pounds containing copper, zinc, lead, gold, and silver. The process is a
geothermal convection cell and it is able to concentrate and segregate useful
rninerals. The process has many points of resemblance to the concentration
and segrega tia n of petroleum - the principal difference being that the geo-
thermal convection cells opérate at higher temperatures.
sio ORIGIN OF OILFIELD WATER

Petroleum and natural gas

Th total amount of organic matter di persed in the sedimentary rocks of


thc carth has been estirnated to be about 2,700 trillion metric tons; of this
amount 50 trillion metric tons are dispersed petroleum hydrocarbons, of
which 0.5 trillion metric tons exist in petroleum reservoirs (Hunt, 1968).
The types of hydrocarbons in a petroleum often indicate its origin; for
example, if it contains predominately odd-numbered n-alkanes in the low
molecular-weight range, it probably was formed from marine organisms.
Petroleums from the Uinta Basin contain a predominance of odd-numbered
hydrocarbons in the wax fraction, indicating a nonmarine organic source.
Waxes derived from land plants contain a predominance of hydrocarbons
with carbon numbers of C27, C29, and C3¡, while hydrocarbons derived from
marine plankton may contain more hydrocarbons with carbon numbers of
CIS' CJ?, and C19•
The water in the sediments containing the organic matter contains many
dissolved organic constituents such as salts of the humic, fatty, and
naphthenic acids, sugars, heterocyclics, and aromatic oxygen compounds.
Degens et al. (1964) observed that as the salt concentration in petroleum-
associated waters increases, the concentration of dissolved amino acids in-
creases.
Petroleum is generated in organic-rich shales, but the mechanisms whereby
it migrates from the shales and concentrates in porous reservo ir rocks are not
understood. Petroleum precursors leave the shale with the water as the water
is expelled by compaction.
As burial proceeds, pressures and temperatures increase. With increasing
temperature, chemical changes in the solids are accelerated and the organic
matter first generates petroleum, which ultimately is converted to methane
and finally graphite. The clay minerals continue their recrystallization, and
finally metamorphism to slates, phyllites, and schists occurs: These processes
involve a continuing loss of porosity with the release of additional pare
water.
The solubility of petroleum hydrocarbons in water increases with in-
creasing temperature and pressure. However, at ambient temperature and
pressure the solubility in pure water is rather low (McAuliffe, 1969). Water-
wet shale has no permeability to immiscible fluids such as gas or oil, so the
petroleum or petroleum precursors probably do not move as droplets. Peake
and Hodgson (1966) report "accommodations" of specific hydrocarbons in
water up to about 30 ppm. Cartmill and Dickey (1970) found that a
colloidal suspension was able to pass through water-wet sands, but the tiny
droplets coalesced at points where the grain size decreased. Neruchev and
Kovacheva (1965) offered some evidence that the amount of extractable
hydrocarbon decreases in shales for the first few meter s away from the
reservoir rocks, as if removed by some flushing action.
Bruderer (1956) suggested that oil deposits originated from sea water
212 ORTGIN OF OILFIELD WATERS

occurs if the equilibria in the aqueous phase remain constant while the water
moves through the sedimentary rocks. Deposition or accumulation of
hydrocarbons occurs if the equilibria in the aqueous phase shifted, causing
desolubilization or precipitation of the hydrocarbons.
Temperature gradients in sedimentary basins usually are about 1°C per 46
m of depth, and the rate of heat flow to the surface is approximately
1.2 x 10-6 cal cm"? sec-¡ (Birch, 1954). Temperature is believed to be a
primary cause in the conversion of organic matter in rocks to petroleum
(Philippi, 1965), it is also believed that lipids are the majar precursors of
petroleum and that most petroleum is generated by chemical reactions oc-
curring at ternperatures above 115°C.
Nonmarine sources are recognized for many crude oils, in contrast to the
once general belief that such sources are unfavorable for the generation of
petroleum. Perhaps the best known examples in the United States are the
nonmarine sequences in the Eocene of the Uinta Basin in Utah. Other
examples of oil and gas with continental source sediments are basins such as
the Dzungaria, Tsaidam, Tarim, Turfan, Ordos, Pre-Nan Shan, and Sungliao
of China (Meyerhoff, 1970). There is considerable nonmarine Tertiary age
strata in the Coa k Inlet-Kena Basin in coastal Alaska.

TABLE7.IV

Tertiary system - highest concentration of a constituent found, average co ncentration,


and number of samples analyzed

Constituent Concentration (mg/l) Number of samples


• highest average

Lithium 27 4 169
Sodium 103,000 39,000 379
Potassium 1,200 220 176
Rubidium 0.6 0.24 11
Cesium 0.4 0.20 9
Calcium 38,800 2,530 376
Magnesium 5,800 530 368
Strontium 420 130 142
Barium 240 60 140
Boron 450 36 170
Copper 1 0.63 3
Chloride 201,300 64,600 380
Bromide 1,300 85 323
Iodide 35 28 322
Bicarbonate 3,600 560 364
Carbonate 300 75 8
Sulfate 8,400 320 139
Organic acid
as acetic 1,900 140 53
Ammonium 2,700 230 64
COMPOSIDIO OF OlLFIELD WATERS 213

TABLE 7.V

Cretaceous system - highest concentration of a constituenl found, average concentration,


and number of samples analyzed

Constituent Concentration (mg/l) Number of samples

highest average

Lithium 13 4 26
Sodium 88,600 31,000 987
Potassium 580 130 38
Rubidium 0.10 0.10 1
Cesium 0.10 0.10 1
Calcium 37,400 7,000 987
Magnesium 8,000 900 987
Strontium 980 200 39
Barium 670 40 34
Boron 70 20 38
Chloride 187,000 62,000 987
Bromide 1,760 550 173
Iodide 190 25 172
Bicarbonate 1,660 260 864
Sulfate 7,100 280 776
Ammonium 35 23 2

Composition of oilfield waters

To il!ustrate the variety of oilfield waters found in subsurface petroleum-


bearing formations, samples were taken, and analyzed, from formations of
the following ages: Tertiary, Cretaceous, Jurassic, Permian, Pennsylvanian,
Mississippian, Devonian, Silurian, Ordovician, and Cambrian. Insufficient
samples from Triassic age strata were available. These data are given in Tables
7.IV-XIII, and al! of the samples were taken from sedimentary basins in the
United States. Each table gives the highest value found in this study in
milligrams per liter for a given constituent, the average values, and the num-
ber of samples used to estimate the average value. The constituents deter-
mined for most of the brines include lithium, sodium, potassium, rubidium,
cesium, calcium, magnesium, strontium, barium, boron, copper, chloride,
bromide, iodide, bicarbonate, carbonate, sulfate, organic acid as acetic
(actual!y organic acid salts but calculated as acetic acid), and ammonium.
Comparison of the lithium content in the samples from Tertiary age strata
(Table 7.1V) with the lithium content of sea water (Table 7.1), indicates that
the average lithium content of 169 oilfield waters is enriched by a factor of
20. At least one sample of oilfield water contained 27 mg/l of lithium or an
enrichment factor of 135 compared to sea water. Table 7.II indicates that
igneous rocks contain up to 65 ppm of lithium and sedimentarv rcck con-
tain up to 46 ppm of lithium.
214 ORIGIN OF OILFIELD WATERS

Jurassic systern - highest concentration of a constituent found, average concentration,


and number of samples analyzed
--------- --------- - --
Consti tuent Concentrabon (mg/l) Number of samples

highest average
---- ----- ----
Lithium 400 10 80
So dium 120,000 57,300 85
Potassium 900 140 9
Rubidium 0.10 0.10 1
Cesium 0.10 0.10 1
Calcium 56,300 25,800 85
Magnesium 5,200 2,500 84
Strontium 2,080 320 9
Bariurn 50 10 7
Boron 50 13 9
Chloride 210,000 141,000 85
Bromide 6,000 1,200 80
Iodide 40 16 8
Bicarbonate 2,640 140 72
Sulfate 1,480 210 78
Organic acid
as acetic 12 12 1

TABLE 7.VII

Permian system - highest concentration of a constituent found, average concentration,


and number of samples analyzed
-- - --
Constituent Concentration (mg/l) Number of samples

highest average
---
Lithium 6 3 3
Sodium 109,000 47,000 54
Potassium 405 170 3
Rubidium 2 0.80 3
Cesium 0.20 0.13 3
Calcium 22,800 8,600 54
Magnesium 5,800 2,000 53
Strontium 10 7 3
Boron 20 8 3
Copper 0.88 0.88 1
Chloride 177,000 92,700 54
Bromide 68 46 3
Io dide 3 3 1
Bicarbonate 281 77 49
Carbonate 36 36 1
Sulfate 3,400 730 41
Organic a cid
as acetic 220 170 2
Arnmon ium 24 24 3
----
COMPOSITION OF OILFIELD WATERS 215

TABLE 7.VIII

Pennsylvanian system - highest concentration of a constituent found, average concentra-


tion, and number of samples analyzed

Constitue nt Concentration (rng/l ) Number of samples

highest average
- -- - -~--- ------------
Lithium 35 7 45
Sodium 101,000 43,000 951
Potassium 710 170 57
Rubidium 2.30 0.55 25
Cesium 8.50 0.15 19
Calcium 205,000 9,100 950
Magnesium 15,000 1,900 947
Strontium 4,500 600 70
Barium 640 30 41
Boron 70 15 54
Manganese 105 60 2
Chloride 270,000 87,600 950
Bromide 3,900 490 57
Iodide 1,410 210 52
Bicarbonate 1,200 130 897
Carbonate 70 40 2
Sulfate 5,400 430 756
Organic acid
as acetic 2,300 . 430 44
Ammonium 3,300 300 51

Table 7.V indicates that oilfield water samples taken frorn Cretaceous age
strata were enriched in lithium with respect to sea-water. The highest lithium
concentration found in 26 samples was 13 mg/I.
Table 7.VI indicates that oilfield waters taken from Jurassic age strata
contain up to 400 mg/I of lithium, which, compared with sea water (Table
7.1), represents a concentration factor of 2,000. Compared to the hydro-
lyzates in sedimentary rocks (Table 7.11), the concentration factor is about
9.
The lithium concentration in oilfield water s taken from Permian age strata
averaged 3 mg/l (Table 7. VII). Only three samples were available for use in
determining this average.
Table 7. VIII indicates that the lithium concentration averaged 7 mg/l in
45 samples taken from Pennsylvanian age strata. This represents a concen-
tration factor of 35 compared with sea water (Table 7.1).
Table 7.IX indicates that the lithium concentration in oilfield waters
taken from Mississippian age strata is enriched by a factor of 45 compared
with sea water. Table 7.X indicates a similar enrichment factor of 250 for
oilfield waters taken from Devonian age strata. For waters from Silurian age
strata (Table 7.XI) the enrichmen t factor found was 185; for the Ordovician
216 ORIGIN OF OILFIELD WATERS

TABLE7.IX

Mississippian system - highest concentration of a constitu~nt found, average concentra-


tion, and number of samples analyzed

Constituent Concentratian (mg/I) Number of samples

highest average

Lithium 55 9 81
Sodium ll5,800 41,500 210
Potassium 5,000 430 80
Rubidium 5 1 47
Cesium 2 0.40 37
Calcium 37,800 8,900 209
Magnesium II ,200 1,600 202
Strontium 3,390 630 52
Barium 20 5 44
Boran 240 40 86
Copper 3 3 2
Manganese 36 12 5
Chloride 206,000 85,000 210
Bromide 1,800 410 88
Iodide 620 llO 89
Bicarbanate 1,590 185 198
Carbonate 450 450 1
Sulfate 3,500 540 191
Organic acid
as acetic 3,070 370 84
Ammonium 700 210 83

age (Table 7.XII), it was 100; and for the Cambrian age (Table 7.XIII), it was
85.
The data in Tables 7.VI-XIII indicate that waters taken from sediments
that formed during the various geologic ages do not a11have the same chemi-
cal composition and that the waters have evolved considerably in comparison
to modern sea water composition (Table 7.I). The manner whereby this
evolution occurred is not completely understood; however, recent studies
have shed some light on the problern. Note the amount of organic acid as
acetic found in waters taken from the sedimentary rocks (Tables 7.VI-XIII).
The organic acids are present in the oilfield waters as organic acid salts.
These organic compounds possibly are a precursor of petroleum and serve as
a transportation mechanism for migration. The exact composition of each
organic acid salt has not been determined. Knowledge of the composition of
these organic acid salts would aid in geochemical studies of petroleum.
Rittenhouse et aL (1969) studied the minor elements in 823 oilfield-water
samples taken from subsurface formations in the United States and Canada.
The data that they found are shown in Table 7.xIV as 25% quartiles, median
concentrations, and 75% quartiles. The dissolved solids are given in grams per
COMPOSITION OF OILFIELD WATERS 217

TABLE 7.X

Devonian system - highest concentration of a constituent found, average concentration,


and number of samples analyzed

Constituent Concentration (mg/l) Number of samples

highest average

Lithium 170 50 29
Sodium 101,000 48,000 85
Potassium 11 ,600 3,100 30
Rubidium 11 4 12
Cesium 1.4 0.5 11
Calcium 129,000 18,000 85
Magnesiurn 26,000 2,900 82
Strontium 2,300 1,000 8
Barium 120 40 7
Boron 90 30 30
Copper 2 2 1
Manganese 200 175 2
Chloride 259,000 115,000 85
Bromide 3,500 1,060 32
Iodide 120 30 32
Bicarbonate 1,000 155 67
Carbonate 60 30 2
Sulfate 1,700 450 74
Organic acid
as acetic 670 130 27
Ammonium 560 110 32

liter, the data followed by pare given in parts per billion (ppb), and the
other data are given in parts per million. They analyzed samples from several
basins as illustrated in Table 7.XIV, and the elements analyzed included
lithium, magnesium, manganese, nickel, cobalt, chromium, copper, potas-
sium, tin, strontium, titanium, vanadium, and zirconium.
Rittenhouse et al. (1969) concluded that elements in oilfield waters com-
monly are present in the following concentrations:

% Na, el
%or ppm ea, S04
> 100 ppm K, Sr
1-100 ppm Al, B, Ba, Fe, Li
ppb (most oilfield waters) c-, Cu, Mn, Ni, Sn, Ti, Zr
ppb (some oilfield waters) Be, co, Ga, Ge, Pb, V, W, Zn

They found no relationship between the constituents in the brine and the
minerals in the aquifer rocks except for potassium. They postulated that
exchange reactions occurred between the clays in the rocks and potassium in
the water to control the dissolved potasssium.
215 ORIGIN OF OILFIELD WATERS

TABLE 7.XI

Silurian syst ern - highest concentration of a constituent found, average concentration,


and number of samples analyzed

Constituent Concentration (mg/l) Number of samples

highest average

Lithium 90 37 8
Sodium 89,000 49,100 14
Potassium 8,400 1,900 11
Rubidium 8 4 2
Cesium 0.4 0.4 2
Calcium 41,000 21,000 14
Magnesium 12,000 4,300 12
Strontium 880 730 2
Barium 15 15 1
Boron 90 30 10
Chloride 195,000 122,000 14
Bromide 1,700 520 11
Iodide 30 17 10
Bicarbonate 270 115 11
Sulfate 3,500 830 13
Organic acid
as acetic 220 90 9
Ammonium 200 80 10

TABLE 7.XII

Ordovician system - highest concentration of a constituent found, average concentration,


and number of samples analyzed

Constituent Concentration (mg/l) Number of samples

highest average

Lithium 70 20 15
Sodium 89,100 31,000 609
Potassium 2,890 990 15
Rubidium 6 2 11
Cesium 0.5 0.2 9
Calcium 39,000 6,100 609
Magnesium 10,900 1,300 607
Strontium 900 340 12
Barium 10 6 10
Boron 80 20 18
Manganese 56 56 1
Chloride 205,600 62,000 609
Bromide 720 300 19
Iodide 70 25 16
Bicarbonate 2,260 270 598
Carbonate 60 25 26
Sulfate 7,600 1,070 583
Organic acid
as acetic 3,300 520 14
Ammonium 630 140 16
RESEARCH STUDIES 219

TABLE 7.XIlI

Cambrian system - high est concentration of a constituent Found , average concentration,


and number of samples analyzed
----- ---- -
Consti tuent Concentration (mg/l) Number of samples

highest aver.age
---- --~_. --
Lithium 40 17 8
Sodium 43,000 23,400 23
Potassium 2,000 440 10
Rubidium 3.3 3.3 1
Cesium 0.6 0.6 1
Calcium 14,500 4,000 23
Magnesium 8,800 1,300 22
Strontium 360 125 7
Boron 13 7 5
Chloride 95,000 46,100 23
Bromide 1,170 520 5
Iodide 40 18 3
Bicarbonate 790 260 23
Sulfate 2,600 1,170 22
Organic acid
as acetic 50 30 3
Ammonium 120 60 3

Compared with sea water the 823 brines were enriched in manganese,
lithium, chromium, and strontium, and depleted in tin, nickel, magnesium,
and potassium. Generally the silicon content varied inversely with the dis-
solved solids content. This agrees with a study of the solubilities of silicate
minerals where Collins (1969b) found that in general the silicon solubilities
decreased with increasing concentrations of dissolved salts at ambient con-
ditions.

Research studies related to the origin of oilfield brines

Tables 7.IV-XIV indicate that the compositions of oilfield brines are not
consistent, and that they are not formed by the simple evaporation or dilu-
tion of sea water. Oilfield brines are found in deep formations that some-
times contain fresher water nearer surface outcrop areas, in formations con-
taining evaporites or in close proximity to soluble minerals, and in forma-
tions close to surface saline waters.
The amounts and ratios of the constituents dissolved in oilfield waters are
dependent upon the origin of the water and what has occurred to the water
since entering the subsurface environment. For example, some subsurface
waters found in deep sediments were trapped during sedimentation, while
other subsurface waters have infiltrated frorn the surface through outcrops.
t-.:>
t-.:>
o

TABLE 7.XIV

Minor elements in 823 oilfield brine samples in United States and Canada * 1
--- ---

Number Lithium Magnesium Manganese Nickel Tin


of -
samples q25 md q75 q25 md q75 q25 md q75 q25 md q75 q25 md q75

I1linois Basin 22 10 15 25 3,000 6,000 8,000 80p 175p 750p NO NO NO NO < 1p 5p


Louisiana and Texas
Gulf Coast 79 NO NO 4 15 250 550 800p 3.5 >5,000p <ip < 1p < 3p NO < 1p 1p
East Texas 88 NO NO NO 150 250 800 1,800p 3.3 >5,000p NO < 1p 3p NO 3p IIp
North Texas 24 NO NO 15 3,000 5,000 6,000 25 45 90 2p 15p 150p NO 12p 25p
West Texas and New
Mexico 148 3 15 25 500 1,000 1,650 200p 1.8 >5,000p <Lp < 1p < 1p NO < 1p 3p
Permian only 74 2 10 25 500 1,000 2,000 180p 1.7 >5,000p <Lp < 1p 3p NO < 1p 3p
Pennsylvanian only 34 3 10 20 500 1,000 1,500 500p 2.8 >5,000p <lp < 1p < 1p <Ip < 1p 3p
Silurian and Devonian
only 15 4 10 25 200 400 560 30p 300p >5,000p <lp < 1p < 1p NO 1p 5p
Ordovician and Cam brian
only 21 10 15 25 500 800 1,000 150p 400p >5,000p <lp < 1p < 1p <1 1p 4p
Anadarko Basin * 3 118 NO 10 35 900 1,550 3,000 600p 5.6 >5,000p <lp 6p 15p NO 2p 4p
Williston Basin, post-
Paleozoic 25 NO NO 10 10 250 2,000 90p 300p 450p >3p < 3p < 3p NO < 1p . < 1p O
;:o
Williston Basin,
Paleozoic 55 18 35 50 300 600 2,000 200p 660p 1,200p NO NO < 3p NO < 1p < 1p O
Powder River Basin 22 NO NO 2 10 40 225 300p 450p 2,000p NO < 3p < 3p NO < 1p < 1p Z
Other Wyoming 28 NO NO 45 20 100 200 60p 300p 1,000p NO NO 3p NO < lp < 1p O
Colorado 18 NO NO NO < 10 30 300 90p 300p 750p <3p < 3p < 3p NO <i l Op <10p "l

California 116 NO NO NO 35 90 175 300p 950p 2,800p 1p 10p 35p NO 2.5p 4.5p O
Sea Water 0.1 1,272 1p-10p 5.4p 3p r:;
Estimated detection
limit
-

-
2 10 1p 1p 1p -
"l

trl
r-
tJ
*1 Medium (md - Rittenhouse et al., 1969) and quartile (q) concentrations in each area; NO = below detection Iimits; p = concentration in ppb, otherwise ppm.
*2 No data, less sensitive methods of analysis used. ~
;¡:.
*3 Inc\udes Oklahoma Platform and Ardmore Basin.
~
trl
;:o
CJJ
::o
trI
m
trI
>-
::o
TABLE 7.XIV (continuedi=' CJ
:r:
m
Number Dissolvedsolids(gil) Cobalt Chrornium Copper Potassiurn ~
of ---- e
U
samples q25 md q75 q25 md q75 '125 md q75 q25 md q75 q25 md q75 S;;
m
Illinois Basin 22 70 98 119 NO ND NO ND 2p 3p <IOp 10p 75p 180 300 400
Louisiana and Texas Gulf
Coast 79 30 69 131 NO NO <5p ND < 1p < 2p <25p <25p <25p 160 300 ~OO
East Texas 88 27 66 116 ND*' NO*' ND*' ND ND 2p < 1p < 1 <1 NO <50 300
North Texas 24 173 222 241 ND NO ND -Cl p < 1p 35p 25p 150p 450p ND 300 1,000
West Texas and New
Mexico 148 61 111 173 ND NO ND '<2p 2p 4p ND lp 10p 120 350 500
Permian only 74 70 143 215 ND NO ND 1p 2p 4p ND 2p 10p 160 400 750
Pennsyvanian only 34 80 115 168 ND ND ND 1p 3p 4p ND < 1p 5p 200 300 400
Silurian and Devonian
only 15 42 55 72 ND ND ND <2p 2p 4p < lp 4p 15p 170 300 450
Ordovie ian and
Carnbrianonly 21 53 67 128 NO NO <5p <í l p < 2p 3p ND. 4p I5p 200 400 650
Anadarko Basin*3 118 51 137 203 NO NO ND -Cl p 10p 25p < lp 10p 25p 20 250 500
Willislon Basin. post-
Paleozoic 25 9 59 88 ND <5p <5p <2p' < 2p < 2p <25p <25p 20p 200 300 400
Williston Basin,
Paleozoic 55 115 173 296 ND ND ND ND 3p 15p ND 3p 5p .400 800 <5,000
Powder River Bas in 22 3 5 11 <5p <5p <5p <2p < 2p < 6p <25p <25p 70p 200 300 400
Other Wyoming 28 4 5 11 NO ND ND ND ND 2p ND ND 3p NO 300 700
Colorado 18 3 5 15 ND <Ep <5p ND ND < 6p <25p <25p <25p 200 300 400
California 116 5 18 30 ND NO 2p 2p Sp 15p 2p 5p 20p NO 45 70
Sea water 35 0.27p 0.04p-0.07p 1p-15p 380
Estimated detection
limil lp Ip lp 50
- ._._-

* ~ Mediu m (md - Rittenhouse el al.,1969) and quartile (q) concenirations in each area; ND = below detectio
n limits; P = concentration in ppb, ot.her
wise ppm.
* No data, less sensitivemethods of analysisused.
*3 Includes Oklahoma Platform and Ardmore Basin.

~
~
>-'
tv
tv
tv

TABLE 7 XIV (con/¡nued)*'

Number Strontium Titanium Vanadium Zirconium


of
samples q25 md q75 q25 md q75 q25
--- - ~~md q75 q25 md
------
q75

IIlinois Basin 22 140 300 400 <10p <10p <10p NO NO NO NO <10p <10p
Louisia na and Texas
Gulf Coasl 79 45 85 200 NO <10p -Cl Op NO NO <it p <10p <10p <101'
Easl T'exas 88 75 350 750 NO NO < 1p NO NO NO NO NO NO
North Texas 24 150 450 700 <10p 7p 20p NO NO NO NO <i op 10p
Wesl Texas and New
Mexico 148 75 200 400 NO <10p <10p NO NO <11' NO NO NO
Permian only 74 65 90 300 NO <10p <101' NO <it p <lp NO NO <10p
Pennsylvanian only 34 180 300 450 NO <10p <10p NO <lp NO NO NO NO
Silurian and Devonian
only 15 75 90 300 <il üp <10p <lOp NO NO <i l p NO NO NO
Ordovician and
Cambrian only 21 100 250 400 NO <10p <10p NO NO -Cl p NO NO NO
Anadar ko Basin 118 90 300 650 NO <10p <10p NO <lp <11' NO <10p 20p
Willislon Basin,
post-Paleozoíc 25 20 100 200 NO NO <lOp NO <11' <11' NO NO NO
Williston Basin,
Paleozoic 55 50 95 450 <10p <10p 25p NO <lp <lp NO NO <lOp
O
Powder River Basin 22 NO 25 50 <lOp <10p <IOp NO <lp <11' <lOp <101' <lOp'
~
,.....
Other Wyoming 28 10 20 45 NO <lOp <10p NO <lp <il p NO NO NO el
Colorado 18 7 20 60 <lOp <101' <10p NO <lp <lp NO <10p <lOp Z
California 116 NO 10 22 <10p <lOp 7p NO -Cl p 2.5p NO NO NO O
Sea water - 13 present 0.3p NO 'lj
Estirnated delection O
limil - 16 IOp 1p 10p
8
'lj

*I Medium (md ; from Rittenhouse el al., 1969) and quarlile (q ) coricentrations in each atea: ND = below det.ect ion limits; p = concentration
R
1:"'
in ppb, otherwise ppm. O
*~ No data, less sensitive rnethods of analysis used. :E
* lneludes Oklahoma Plat íorm and Ardmore Basin. >-
>-3
tt1
~
tr:
RESEARCH STUDIES 223

Some waters are mixtures of the infiltration water and trapped ancient sea
water.. Also, the rocks containing the waters often contain soluble con-
stituents which dissolve in the water s or contain chemicals which will
exchange with chemicals dissolved in the waters causing alterations of the
dissolved constituents.
The amounts of dissolved constituents found in oilfield waters range fram
less than 10,000 mg/I to more than 350,000 mg/l. This salinity distribution
is dependent upon several factors including hydraulic gradients, depth of
occurrence, distance from outcrops, mobility of the dissolved chemical
elements, soluble material in the associated rocks, ion exchange reactions,
and clay membrane filtration.
Concentration of sea water can occur by surface evaporation, and there
are at least three independent processes that can cause major changes in
buried, isolated sea water:
(1) Dilution with meteoric or fresher waters which have entered outcraps.
(2) Reactions with minerals in the sediments and sedimentary rocks (the
reactions are often temperature and pressure dependent).
(3) Membrane filtration through clays and shales as a result of pressure
and osmosis.

Playa deposits

Jones et al. (1969) studied the composition of brin es in shallow, fine-


grained playa deposits in the Great Basin. The concentrations of dissolved
solids in the water in these sediments often were as much as five times
greater than in the water in the associated lakes. They attributed the concen-
tration pracesses to capillary evaporation and entrapment of fossil brin es
when the salinity of the lake water was greater (lake nearly dry).

Continental Slope drill holes

Manheim and Bischoff (1969) analyzed pore waters from drill holes on
the Continental Slope of the northern Gulf of Mexico. A relationship be-
tween the salinities of the waters and the proximity of diapiric structures
was found. This indicated that salts are leached from salt-bearing sediments
to increase the salinities of the pore waters. In some samples the high
bromide and potassium concentrations suggcsted that late-stage evaporitic
minerals such as carnallite and polyhalite were leached frorn salt bodies.
They postulated that molecular diffusion is a major mechanism which in-
fluences the distribution of salt in the pore waters. Similar conclusions have
been made for saline waters in other areas.

Relation to petroleum accumulations

Van Everdingen (1968) suggested that major circulation svstems of íor-


224 ORIGIN OF OILFIELD WATERS

mation water exist in the Western Cenada Sedimentary Basin, that the flow
systems affected accumulations of hydrocarbons in the basin, and further
that pressure and salinity variations might be explained by membrane
properties of the shales. He saw a need for studies of the hydrodynamics of
the basin.
Hydrodynamics and geochemistry of the Paradox Basin were studied by
Hanshaw and Hill (1969). The ground-water movement in the basin is
generally southwestward frorn the high outcrop areas in western Colorado,
flowing toward the Colorado River discharge areas. Hydrodynamic con-
ditions exist in lower Paleozoic strata which are favorable to accumulations
of petroleum in stratigraphic traps. Paleozoic aquifers in northwestern New
Mexico have very high potentiometric surfaces and these aquifers may be the
outflow receptors of an osmotic membrane system operating within the San
Juan Basin. This regional study was excellent and of value in exploration for
petroleum and gas.
Parker (1969) studied brines and waters in five aquifers of Cretaceous age
in the East Texas Basin. He found that the campo sitian of the waters in the
older, more deeply buried aquifers were modified more than waters in
younger, less deeply buried aquifers. Most of the modifications were made
by exchange reactions, dilution by meteoric waters, and loss of sulfate be-
cause of bacterial reduction. Hydrodynamic movement of the waters in the
Woodbine formation contributed to the giant oil accumulation in the East
Texas Basin. Much of the stratigraphic trapped oil probably was trapped in
part because of this type of flow.

Sabkha sediments and transport of valuable ores

• Bush (1970) discussed the origin of chloride-rich brines from Sabkha sedi-
ments and how they are related to inclusion brines and lead-zinc deposits of
the type found in the Mississippi Valley. He noted that Helgeson (1964) and
Barnes and Czamanske (1967) have shown that chloride-rich solutions can
transport lead and zinc as chloride complexes. According to Bush (1970),
Sabkha brines free of sulfur are expelled by sediment compaction, migrate,
and become enriched in base metals until they contact a zone of higher
temperature and pressure. In this zone, sulfides are present as a result either
of inorganic reduction of sulfate, anaerobic reduction, al' hydrocarbon
reduction of anhydrite. The sulfides cause the base metals to precipitate.
Formation brines are the medium in which several metals, in addition to
hydrocarbons, migrate prior to deposition in ore deposits. A current theory
is that the metal s travel primarily as chloride complexes in solutions that are
depleted in reduced sulfur species (Dunham, 1970). The metals subsequently
are precipitated when a source of reduced sulfur is meto An example of a
source of reduced sulfur is an area where anaerobic bacteria are reducing
sulfate. This occurs in waters near petroleum-bearing formations, and such
waters in carbonate reservoirs often contain considerable amounts of sulfide.
RESEARCH STUDIES 225

Brine classification

A study of the evolution of subsurface brines in Israel by Bentor (1969)


led him to classify brines into four groups. The first group consists of brines
similar to sea water except for an increased concentration of calcium and a
decreased concentration of magnesium, which he attributed to dolomitiza-
tion. The second group was similar to sea water but contained two to three
times higher concentrations of dissolved salts, deficient in sulfate and
magnesium, and enriched in bromide and iodide. The sulfates were lost by
organic reduction, the magnesium was lost by exchange reactions with c1ays,
and bromide and iodide were added by organic sources. The third group was
a high-salinity calcium chloride-type brine formed by surface evaporation
and later modified in the subsurface by differential ultrafiltration. The
fourth group was a highly saline, calcium chloride type with Ca/Na ratios
greater than one. This group was divided into two subgroups where the first
subgroup is a highly saline and highly differentiated Early Paleozoic brine,
while in the second subgroup they are old Paleozoic brin es which were
submitted to an additional cyc1e of surface concentration by evaporation.

Ion association

Truesdell and Jones (1969) studied ion association in brines and found
that, except for the chloride ion, the major simple ions form ion pairs, while
the minor and trace metals in brines form coordination complexes. Selective
ion electrodes can be used to determine directly the ionic activities of
sodium, potassium, chloride, fluoride, and sulfide in brines. Experimental
data were used to calculate chemical models for ion association and coordi-
nation complexes in brines. These models are useful in explaining the
chemical behavior of brines.

Relation to lithology

Kramer (1969) used factor analysis to study the relationships of the brines
to the type of rock from which they were taken. His results indicated that
the major ions in most brines are sodium, calcium, and chloride; brines are
enriched in calcium and bicarbonate and are deficient in magnesium and
sulfate relative to sea water. The factor groupings did not reflect the lithol-
ogy of the rocks from which the brines were taken, indicating that such a
relationship does not exist or is difficult to detecto The brin e analyses used in
the study were primarily macro analyses and did not inc1ude pH, minor, or
trace constituents. A study of this type would benefit significantly if the
following conditions were met: (1) use only the best available sampling
methods; (2) use field analysis techniques; (3) use positive lithology identifi-
cation; and (4) use only the best available laboratory methods of analysis.
226 ORIGIN OF OILFIELD WATERS

Kramer (1969) did 110L have these controis because he used only the
published data of various laboratories.
Carpenter and Miller (1969) used statistical and thermodynamic methods
in an effort to determine the origin of the dissolved chemical constituents in
saline subsurface waters in north-central and northwestern Missouri. Statisti-
cal analysis oí scatter diagrams indicated that the concentrations of lithium,
sodium, potassium, and bromide and the ion activity ratios of K+/H+ ,
Ca" 2 /Mg+2, and Sr+ 2 /Ba+ 2 in the waters are influenced by reactions with
constituents in the aquifer rocks. They concluded that the ion ratios are of
little value in determining the origin of the waters beca use the concen-
trations of the dissolved constituents in the waters had reacted with minerals
in the aquifer rocks. This study was excellent because it did show that the
concentrations of constituents in the water are controlled to some extent by
reactions with the aquifer rocks. Additional work of this kind is needed in
the study of deep brines.
A study of brines from the Sylvania formation in the Michigan Basin
indicated that evaporation and dolomitization were two dominant controls
for their dissolved concentrations of calcium, magnesium, sodium, stron-
tium, and bromide (Egleson and Querio, 1969). Mechanisms responsible for
concentrations of elements such as potassium, lithium, rubidium, ammonia,
boron, and iodide were believed to be reactions with sedimentary racks,
leaching of organic constituents, and bioconcentration.

Relation to depth and salinity

A study of the chemical composition of some selected Kansas brines


indicated that in general the concentrations of calcium, sodium, and chloride
increase with increasing salinity, while the sulfate concentrations decrease
(Dingman and Angino, 19(9). However, the Ca/CI ratio, concentrations of
calcium and salinity, did not generally increase with geologic age or with
depth of the aquifer.
Dickey (1969) surveyed the analyses of oilfield waters frorn many areas of
the United States and concluded that in general the Ca/Mg ratio increases
with increasing salinity while the ratio Na/(Ca + Mg) decreases irregularly
with increasing salinity and depth. This observation is compatible with the
findings of several investigators. The dominant anion in subsurface waters
usually changes with depth; in near-surface waters, it is sulfate; at depths
exceeding 520 m it is bicarbonate; and in deep brines, it is chloride
(Chebotarev, 1964). The Ca/Na ratio usually increases with depth and age of
the associated rocks, while the Mg/Na ratio decreases.

Iodide

Collins (1969a) studied the chemistry of some oilfield brines from the
Anadarko Basin which contain high concentrations of iodide. The concen-
RESEAReH STUDIES 227

trations of bromide in many of these brines are lower than the iodide which
is unusual. Localized sedimentary rock deposits enriched in organic iodine
are the source of the high iodide concentrations in these brines (Collins et
al., 1971).

Hot brines

Hot brines containing minor and trace amounts of several metallic


elements in addition to macro concentrations of some alkalies, alkaline
earths, and chloride are found in drill holes in southern California, in the
Caspian Sea, and in deeps in the Red Sea. These brines were formed from
evaporites dissolved by meteoric water, and the metallic elements were
leached from country rocks by the hot brines (Tooms, 1970). Laboratory
reactions of 2M and 4M sodium chloride with andesite and shale at
0 0
300 -500 C have produced solutions containing metallic elements in con-
centrations similar to the hot brines (Ellis, 1968).

Comparison of oilfield brines with evaporated sea water

Bromide does not form its own minerals 'when sea water evaporates. It
forms an isomorphous admixture with chloride in the precipitates
(Valyashko, 1956; Braitsch and Herrmann, 1963). As sea water evaporates,
the carbonates precipitate first, followed by the sulfates. Little or no
bromide precipitates, or if it does, it is occluded with these.
Halite (NaCl) begins to precipitate when the chloride concentration is
about 275,000 mg/l (Table 7.III) compared with that of normal sea water,
19,000 mg/l. Some bromide is entrained with chloride in the precipitate.

300,------------------------------------------,

200
-- ---
Normal evaporite curve
e J
e
e

:::::: 100
~ e
o- e
-
_TiTT*f~i
w t:
o
a::
o
...J 50 TVT T
T TT T T ee J
e
e
I
u
30
iI"
T
T T
ee e e
Te e J
T T
20 e
e

Fig. 7.4. Use of the bromide ion to differentiate some Tertiary (T), Cretaceous (e), and
Jurassic (J) age brines.
228 ORIGIN OF OILFIELD WATERS

TABLE 7.XV

Bioconcentrated bromide and iodide in sea wee ds and corals

Bromide (ppm) Iodide (ppm)

Seaweed
Laminaria digitata (dry matter) 1,380 510-8,000
Laminaria saccharina (dry matter) 340 2,000
Desmaresta (ash) 6,800 5,200

Corals*
Gorgonia uerrucosa 16,200 69,200
Gorgonellidae 19,800 22,100
Isididae 7,400 20,300

* After Vinogradov (1953).

However, with each crystallization, more bromide is left in solution than is


entrained in the precipitate.
Sylvite (KCI) begins to precipitate when the chloride concentration is
about 360,000 mg/l (Table 7.I1I), followed by carnallite (MgCI2 "KCl "6H20)
and bischoffite (MgCI2 "6H2 O). During evaporation the concentration rate of
bromide in solution increases. The change in the slope of the curve in Fig.
7.4 illustrates this approxima tely.
Other concentration mechanisms operate to account for the high bromide
concentrations (6,000 mg/l" and up) found in some brines. One of the
mechanisms is related to bioconcentrators such as seaweeds and corals. The
seaweeds and corals concentrate the bromide, they die, and are buried with
the sediments. Later the bromide is leached by the surrounding waters. Table
7.XV illustrates some of the concentrations of bromide and iodide that
Vinogradov (1953) found in various seaweeds and corals.
Laboratory experiments have demonstrated that bromide is accommo-
dated in the halite crystal lattice and replaces chloride in solid solution
(Borchert and Muir, 1964). The weight percentage of bromide in solid solu-
tion in the halite lattice is related to its weight percentage in the parent brine
as:
wt.% Br (in halite)
e = wt.% Br (in solution)
where e = the partition coefficient.
In most natural environments e = 0.14 (Braitsch and Herrmann, 1964).
In a marine salt sequence the wt.% Br/NaCI rises from about 0.007 wt.% at
the bottom to 0.02 wt.% at the beginning of potassium precipitation. How-
ever, the bromide concentration with a given natural halite sequence may
vary considerably, even though theoretically it should in crease continuously
RESEARCH STUDIES 229

from the bottom to the top of the depositional strata. These variations can
be attributed to inflow of fresh sea water during the deposition or subsequent
leaching after depositíon. Rittenhouse (1967) developed a method to classify
oilfield water s based upon the bromide concentrations.
Fig. 7.4 is a log-log plot of chloride versus bromide concentrations for
some Louisiana oilfield waters. The T, C, and J on the figure refer to Terti-
ary, Cretaceous, and Jurassic, indicating the ages of the rocks from which the
waters were taken. The normal evaporite curve was plotted by using data
from Table 7.III. The data in the figure indicate that most of the Tertiary
waters are deficient in bromide when compared to an evaporite water,
whereas the Cretaceous and Jurassic waters are enriched in bromide (Collins,
1967).
The Tertiary waters contain dissolved halite, which accounts for their low
bromide concentration, while the waters that are enriched in bromide con-
tain bitterns or have leached bromide from sediments that were enriched in
bioconcentrated bromide.
The bromide content of oilfield brines can be used to distinguish between
brines that originated because .of evaporation of sea water and those formed
by the dissolution of evaporite minerals. This can be done by using Fig. 7.4.
If the bromide concentration falls to the right of the normal evaporite curve,
the brine contains evaporated sea water, while if it falls to the left of the
curve, it contains dissolved evaporite minerals.
Fig.7.5 illustrates how closely the concentration of sodium of some
Louisiana oilfield waters taken from Tertiary, Cretaceous, and Jurassic age
rocks follow the sodium concentration of a brine associated with normal
evaporation (Collins, 1970).

300 .~,-------------------------------------.

200
---
_ 100
<,

'"
w
o
a: 50
o
..J

e 30

20

10 5,000 10,000 20,000 50,000 100,000


SODIUM, mg/I

Fi¡t. 7.5. R lationships of the chloride concentrations to sodium concentrations in a


normal evaporite brine to oilfield brilles taken from rUl1l1aLiom. ~f TCl'tilU!'y (T), ('YQ.
taceous (C), and Jurassic (J) age in the United States.
230 ORIGIN OF OILFIELD WATERS

Jon excnang»

Ion exchange reactions on clay minerals are reversible and they follow the
law afmass action. The number of exchange sites governs the reaction, and
other important factars include temperature, pressure, solution cancen-
trations, and banding strength of exchangeable ions. Ion exchange between
clay minerals and a brine will stop when equilibrium is attained.
As the waters move in their subsurface environment, their dissalved ions
have a tendency to exchange with those in the rocks. There are two extreme
types of adsorption in addition to intermediate types of adsorption. The
extreme types are: (1) a physical adsorption or Van der Waals adsorption
with weak banding between the adsarbent and the constituent adsorbed; and
(2) a chemical adsorption with strang valence bonds.
Cations can be fixed at the surface and in the interior of minerals. These
fixed cations can exchange with cations in the water. Under the right
physical conditions of the adsorbent, similar exchange can occur with the
anions, Sorne of the canstituents in f'ormations that are capable of exchange
and adsorption are argillaceaus minerals, zeolites, ferric hydroxide, and cer-
tain organic compounds.
Particle size influences the exchange rates and capacities if the solids are
clays such as illite and kaolinite. The rate increases with decreasing particle
size. However, if a larger mineral has a lattice, the exchange can easily occur
an the plates. The concentration of exchangeable ions in the adsorbent and
in the water is important. More exchange usually occurs when the solution is
highly cancentrated.
According ta Grim (1952), the replacing power of same ions in clays is:
(1) In NH4, kaolinite:
Cs > Rb > K > Ba > Sr > Ca > Mg > H > Na > Li
(2) In NH4, montmorillonite:
Cs> Rb > K > H > Sr > Ba > Mg > Ca > Na > Li

These two clays aften are present in sedimentary rocks and the replacing
arder indicates that lithium and sodium are more likely to be left in solution,
while cesium and rubidium are mare likely ta be removed fram solution.
Fig. 7.6 is a plot af the chloride content versus the lithium cantent of
sorne oilfield water s taken frorn the Smackaver forrnatíon. The .lithium
enrichment results at least in part frorn exchange reactians on clays. Lithium
has a small radius, a low atornic number, a larger hydrated radius than
sodium, and a larger polarization than sodium. Because of these, its replacing
power in the lattices of clay minerals is low (Kelley, 1948). Other ions such
as barium, strontium, calcium, magnesium, cesium, rubidium, potassium, and
sodium will preferentially replace lithium in clay minerals, thus releasing
lithium to solutions. Furthermare, the solubility products af most lithium
RESEAReH STUDIES 231

1,000,---------------------_--,
800
o Louisiono
600 • Mississippi
• Alabama
400
o Arkansas
o Texas
200 , •• : A 00
;;:: 'l,o

W
'" -Norrno! evaporite
curve o o
00 00

O 100
ir 80
O
...J
I 60
U

40

20

10
2 4 6 10 20 40
LlTHIUM, mg/I

Fig. 7.6 Relationships of the concentrations of chloride and lithium in a normal


evaporite-formed brine to oilfield brines taken from the Smackover formation in five
states of the United States.

300,----------------------------,

200
Norma 1 evoporile curve----.
C

_ 100 cc ro:
J J

: lT ~ TT~tij(~T i
~ 50 J C J ~1f~'t
T TT T
...J Tc¡: C T C T C
G 30 .~CC
T
20 C-
J
T

500 1,0002,000
POTASSIUM, mg/ 1
Fig. 7.7. Relationships of the concentrations of chloride and potassium in a normal
evaporite-formed brine to oilfield brines taken frorn formations of Tertiary (T), Cret a-
ceous (e), and Jurassic (J) age in the United States.

compounds are higher than those of other alkalies and alkaline earths. There-
fore lithium tends to stay in solution.
Fig. 7.7 compares tho potassium eoncsntration of soma Louisiana oilfield
waters with those of waters subjected to evaporation. All of these waters are
depleted in potassium with respect to a brine subjected to evaporatian
indicating that potassium was lost to the assaciated sediments during
232 ORIGIN OF OILFIELD WATERS

diagenesis. It has been shown that a tendency exists for potassium to be


adsorbed and fixed by clay minerals, mica, and potassium feldspar in normal
low-temperature processes (White, 1965; Khitarov and Pugin, 1966; Grim,
1952).
The data in Tables 7.III- XIV indicate that the concentration of calcium
in oilfield waters generally is enriched relative to sea water. Cation exchange
reactions with clays accounts for some of this enrichment:

2Na+ (solution) + Ca (clay) -+ Ca+2 (solution) + 2Na (clay)

Collins (1972) found that the ratio Na/(Cú + Mg) tends to decrease as the
dissolved solids concentration increases in some oilfield waters from the East
Texas Basin. This depletion of sodium with respect to calcium plus magne-
sium was attributed to diagenesis of the waters and it correlated with an
index of base exchange (Schoeller, 1955), indicating that the alkali metals in
the waters exchanged with alkaline earth metals on the argillaceous minerals
to decrease the dissolved alkali metals and increase the dissolved alkaline
earth metals.
Fig. 7.8 is a plot of the calcium concentration in some oilfield waters
taken from the Smackover formation. All of these waters are enriched in
calcium relative to the evaporated sea water.
Krejci-Graf (1963) found that solutions predominantly concentrated in
chloride can force an exchange of calcium and bromide from clay minerals
for sodium and chloride from the solution. If this type of reaction occurred

1,000
800
o Louisiana
600 • Mississippi
• Alabama
400
o Arkansas
o Texos
;;::: 200

W
'" ....-Normal
curve
evaporite

o
t:t: 100
o
..J 80
I
u 60

40

20

10

Fig. 7.8. Relationships of the concentrations of chloride to calcium in an evaporite-


formed brine to oilfield brines taken from the Smackover formation in five states of the
United states.
RESEARCH STUDIES 233

1,000,----------------------,
800
6OO 6 Louisiono
• Mississippi
400 • Alabomo
o Arkonsos
o Texas
- 200
<,

'"
w
o 100
g:¡ 80 o
I! 60 'Normal evoporite
o

u
curve
40

20

100 200 400

Fig. 7.9. Relationships of the concentrations of chloride to bromide in an evaporite-


formed brine to oilfield brines taken frorn the Smackover formation in five states of the
United States.

to the Smackover brines, it explains their enrichment of calcium and


bromide.
Kozin (1960) wrote about a "reverse" exchange of anions when the
cations exchange on clays:

CI- (solution) + Br (clay) -+ Br" (solution) + CI (clay)

Such a reaction al so helps to account for the bromide enrichment found in


most oilfield waters taken from the Smackover formation (Fig. 7.9).
A similar reaction for iodide:

CI- (solution) + 1 (clay) -+ 1 (solution) + CI (clay)

would help explain the tremendous enrichment of iodide in oilfield brines


(Collins, 1969a) with respect to sea water as demonstrated in Tables
7.1V-XIII.
Fig. 7.10 shows that boron usually is enriched relative to the normal evap-
orite curve in Smackover oilfield brines. Boron, like lithium, has a small
radius, a low atomic number, and large polarization. Therefore, its replacing
power in the lattices of clay minerals is low. Also, boron does not have a
tendency to enter silicate lattices of the common rock-forming minerals.
Because of these factors, it usually remains in solution until late-stage crys-
tallization.
234 ORIGIN OF OILFIELD WATERS

1,000 r------------------------,
800
600 c> Louisiana
• Mississippi
400 • Alabama
o Arkansas
o Texas
200

o o
w
ºg¡ 100
80 'Normal evaporite
~ 60 curve
u
40

20

10~LLLUU_~-L_~~JJ~L_~~~~~LLuu_~
4 6 10 20 40 60 100 200 400 1,000 2,000
BORON, mg/I

Fig. 7.10. Relationships of the concentrations of chloride to boron in an evaporite-


formed brine to oilfield brin es taken frorn the Smackover formation in five states of the
United States.

Mineral formation

A study of some brines taken from Devonian age reservoir rocks indicated
that dolomitization probably is tlie most important mechanism in deter-
mining the calcium, strontium, and magnesium content of these brines
(Egleson and Querio, 1969). It also was concluded that the relative amounts
of ammonium, iodide, and lithium in these brines were too high to be
derived directly from sea water, and the ammonium and iodide probably
were enriched in the brines as a result of bioconcentration and subsequent
leaching of organic debris.
Dolostone deposits owe their origin to hypersaline brines (Friedman and
Sanders, 1967). Some dolornite, including diagenetic and epigenetic forms,
originates from subsurface brines. In the geologic columns in several oil-

TABLE 7.XVI

Approximate sea-water composition before and after gypsum precipitation (mg/l)

Ion Before precipitation After precipitation

Calcium 390 o
Magnesium 1,300 1,300
Bromide 65 65
Sulfate 2,580 2,580
RESEARCH STUDIES 235

TABLE 7.XVIl

Approximate sea-water composition after dolomitization or bacterial reduction (mg/l)

Ion After dolomitization After bacterial reduction

Calcium o O
Magnesium 883 1,300
Bromide 65 65
Sulfate O O

productive basins, mixtures of dolomite and anhydrite occur, which in-


dicates that sulfate may have been removed from the associated waters by
dolomitization as well as by bacterial reduction.
Table 7.XVI illustrates the approximate amounts of calcium, magnesium,
bromide, and sulfate that could exist in a water ·before and after precipita-
tion of gypsum.
Assuming that the residual sulfate (1,644 mg/l) was removed by the
dolomitization reaction:

MgCl2 + 2CaC03 ~ CaCl2 . + CaMg(C03 h


CaCl2 + MgS04 ~ CaS04 + MgCl2
MgS04 + 2CaC03 ~ CaS04 + CaMg(C03 h

then the Mg/Br ratio would be about 883/65 = 13.6, as illustrated by the
data in Table 7.XVII. However, if the residual sulfate was removed by bacte-
rial reduction:

the Mg/Br ratio would be about 1300/65 = 20.


Magnesium will react with CaC03 (calcite) to form dolomite, thus in-
creasing the concentration of calcium in the brine. However, the total cal-
cium plus magnesium in the brine should remain constant. This can be
calculated as (24.31/40.08) x mg/I calcium + mg/l magnesium = total equiv-
alent magnesium or Mg'. The ratio Mg' /Mg will vary, depending upon the
availability of calcite, and the ratio should be indicative of the degree of
dolomi tiza tion.
For example, brines that are in equilibrium with sandstones should have a
relatively low Mg' /Mg ratio, those in equilibrium with dolomite should have
higher ratios, and those in equilibrium with limestone should have the
highest ratios. The average ratio for some Smackover brin es is 7 (Table
7.XVrII), which indicate that the brines wsre in equilibrium with limestone
and dolomite. Brines from so me Tertiary age rocks which were primarily
236 ORIGIN OF OILFIELD WATERS

TABLE 7.XVIII

Concentration ratios and excess factor ratios for some constituents in Smackover brines

Constituent Average composition (mg/l) Concentration Excess Number of


ratio* 1 factor*2 Smackover
sea water Smackover samples
brines

Lithium 0.2 174 870 18.1 71


Sodium 10,600 66,975 6 0.1 283
Potassium 380 2,841 8 0.2 82
Calcium 400 34,534 86 1.8 284
Magnesium 1,300 3,465 3 0.1 280
Strontium 8 1,924 241 5 85
Barium 0.03 23 767 16 73
Boron 4.8 134 28 0.6 71
Copper 0.003 1.1 359 7.5 64
Iron 0.01 41 4,049 84.2 90
Manganese 0.002 30 14,957 311 69
Chloride 19,000 171,686 9 0.2 284
Bromide 65 3,126 48 1 74
Iodide 0.05 25 501 10.4 73
Sulfate 2,690 446 0.2 0.003 271
Mg'*3 1,543 24,362 16 0.3 284

* 1 Amount in brine/amount in sea water.


*2 Concentration ratio of a given constituent/concentration of bromide.
*3 Mg' = (24.31/40.08) x mg/l Ca + mg/l Mg.

sandstone (Table 7.XIX) had an average ratio of 2.8, while brines from some
Cretaceous age rocks had an average ratio of 6.0 (Table 7.XX).
Bromide does not form its own minerals when sea water evaporates. Some
of it is lost from solution because it forms an isomorphous admixture with
chloride with the halite precipitate. However, more bromide is left in solu-
tion than is entrained in the precipitate. Therefore, relative to chloride, the
bromide concentration in the brine increases exponentially. Because of this,
the bromide concentration in the brine is a good indicator of the degree of
sea water concentration, assuming that appreciable quantities of biogenic
bromide have not been introduced.
Table 7.XVIII presents data that were obtained by comparing the average
composition of some Smackover brines with that of sea water. The concen-
tration ratio was calculated by taking the mean average for a given con-
stituent in the Smackover brines and dividing it by the amount of the con-
stituent found in normal sea water. The excess factor was determined by
dividing the concentration ratio of a constituent by the concentration ratio
of brornide. The calculation for Mg' or total equivalent magnesium was
previously explained, and the number of Smackover samples indicates how
RESEARCH STUDIES 237

TABLE 7.XIX

Concentration ratios of some constituents in some brines taken from Tertiary age rocks

Constituent Average composition (mg/l) Concentration Excess


ratio*l factor * 2
sea water Tertiary brines

Lithiurn 0.2 3 15 12.5


Sodium 10,600 37,539 3.5 2.9
Potassium 380 226 0.6 0.5
Calcium 400 2,077 5.2 4.3
Magnesium 1,300 686 0.5 0.4
Strontium 8 148 18.6 15.5
Barium 0.03 73 2,439 2,033
Boron 4.8 20 4.1 3.4
Chloride 19,000 63,992 3.4 2.8
Bromide 65 79 1.2 1
Iodide 0.05 21 426 355
Sulfate 2,690 104 ü.03 0.03
Mg' 1,543 1,947 1.3 1.1

* 1 Amount in brine/amount in sea water.


*2 Concentration ratio of a given constituent/concentration of bromide.
*3 Magnesium equivalent of caJcium plus magnesium in brine: Mg' = (24.31/40.08) x mg/l
Ca + mg/l Mg'.

TABLE 7.XX

Concentration ratios of some constituents in some brines taken from Cretaceous age rocks

Constituent Average concentration (mg/l) Concentra tion Excess


ratio* 1 factor*2
sea water Cretaceous brines

Lithium 0.2 4 20 4.5


Sodium 10,600 28,462 2.7 0.6
Potassium 380 193 0.5 0.1
Calcium 400 4,999 12.5 2.8
Magnesium 1,300 606 0.5 0.1
Strontium 8 346 43.3 9.8
Barium 0.03 48.3 1,608 365
Boron 4.8 27.5 5.7 1.3
Chloride 19,000 54,910 2.9 0.7
Bromide 65 287 4.4 1
Iodide 0.05 37 737 168
Sulfate 2,690 206 0.1 0.02
Mg'*3 1,543 3,643 2.4 0.5

* 1 Amount in brine/amount in sea water.


*2 Concentration ratio of a given constituent/concentration of bromide.
*3 Magnesium equivalent of calcium plus magnesium in brine: Mg' = (24.31/40.08) x mg/l
Ca + mg/l Mg' ..
238 ORIGIN OF OILFIELD WATERS

many samples were used in the calculation. For example, 71 Smackover


brines were analyzed [01' lithium, while 283 were analyzed Ior sodium.
The concentration ratios (Table 7.XVIII) indicate that all of the deter-
mined constituents in the Smackover brin es were enriched with respect to
sea water except sulfate. However, the excess factor ratios indicate that
sodium, potassium, magnesium, chloride, sulfate, and total equivalent mag-
nesium were depleted in the Smackover brines, while lithium, calcium, stron-
tium, barium, copper, iron, manganese, and iodide were enriched. Further,
these ratios indicate that the Smackover brines have been altered consider-
ably if it is assumed that they originally were sea water.
The concentration ratio 48 for bromide (Table 7.XVIII) is one of the
highest that this author has seen. For example, bromide concentration ratios
of 1.2 (Table 7.XIX), 4.4 (Table 7.XX), and 8.8 and 7.2 were found for brin es
from Tertiary, Cretaceous, Pennsylvanian, and Mississippian age rocks
(Collins, 1967, 1969a, 1970). The concentration ratios and excess factors in
Tables 7.XIX and XX indicate several constituents are enriched and several
are depleted in these brines also.
Almost one-third of the magnesium in sea water and subsequent bitterns
can be removed during the dolomitization reaction. The formation of
chlorite from montmorillonite requires about 9.2 moles of MgO per mole of
chlorite (Eckhardt, 1958):

1.7 Al203 • 0.9 MgO' 8 Si02 • 2 H20 + 9.2 MgO + 6 H20 ~

10.1 MgO . 1.7 Al203 • 6.4 Si02 • 8 H20 + 1.6 Si02

1,000
800
600 ..
o Louisiono
Mississippi
Alobomo
400
o Arkonsos

•.
o Texos

:::::: 200
o
o ~ ~ .0 8'" o
'"
W
800
o
o
o
ir 100
o
--'
80 -; Normol evoporite
curve
I 60
o
40

20

100 200 400 1,000 2,000 4,000 10,000 40,000


MAGNESIUM, mg/I

Fig. 7.11. Relationships of the concentrations of chloride to magnesium in an evaporite-


formed brine to oilfield brines taken from the Smackover formation in five states of the
United States.
RESEARCH STUDIES 239

1,000,--------------------,
800
6 Louisiona
600 • Mississippi
/Normal evaporite • Alabama
400
( curve O Arkansas
O Texas

200

º
o::
o
:r!
100
80
60
u
40

20

Fig. 7.12. Relationships of chloride to strontium in an evaporite-formed brine to oilfield


brines taken from the Smackover formation in five states of the United States.

Such a reaction could remove large amounts of magnesium from waters.


Hiltabrand (1970) has shown that contemporary argillaceous sediments can
remove 100 mgjl of magnesium from sea waters.
Fig. 7.11 is a plot of the chloride concentrations versus the magnesium
concentrations in some oilfield waters taken from the Smackover formation.
The figure indicates that the Smackover waters are depleted in the concen-
tration of rnagnesium with respect to an evaporite-formed brine. Tables
7.111-XIII indicate that in general oilfield waters taken from rocks of other
forrnations also are depleted in magnesium. The data also show that general-
Iy as the dissolved rnagnesium decreases the dissolved calcium increases. This
is related to the forrnation of minerals such as chlorite or dolomite and to
exchange reactions with argillaceous minerals. It is not a result of solubility
because most magnesium compounds are more soluble than calcium com-
pounds.
Pig. 7.12 is a plot of the concentration of chloride versus the concen-
tration of strontium found in some oilfield brin es taken from the Smackover
forrnation. This figure indicates that the strontium concentration is enriched
in the Smackover brines relative to sea water. Reactions that account for
some of this enrichment are:

2SrC03 + MgCl2 ~ SrMg(C03 h + SrCI2


SrMg(C03 h + MgC12 ~ 2MgC03 +SrCI2

The data in Fig. 7_7 and TabJes 7.TH-XIV indicate that the concentration
of potassium in oilfield waters generally is depleted relative to sea water.
Montmorillonite-type minerals systematically change to illite with depth
240 ORIGIN OF OILFIELD WATERS

in Gulf Coast shales (Burst, 1969). As a result of this transformation, the


montmorillonite-type minerals lose interIayer water. Laboratory experiments
at elevated temperatures and pressures indicate that montmorillonite loses its
interIayer water and transforms into illite in the presence of potassium-
enriched water (Khitarov and Pugin, 1966). The structural variations of the
expandable mineral s in clays also are apparentIy infIuenced by the potassium
content of the associated waters. This indicates that oilfield waters tend to
become depleted in potassium content where this reaction occurs.
Reactions between brines and minerals to form silicates that account for
the depletion of dissolved alkali metals are:

3A12 Si2 05 (OH)4 + 2K+ ~ 2KA13Si3 010 (OHb + 3H2 0+ 2H+


KAI3Si030lo(OHb + 6Si02 + 2K+ ~ 3KAlSi30s + 2H+
A12 Si2 OdOH)4 + 4Si02 + 2Na+ ~ 2NaAlSi3 Os + H2 0+ 2H+

These reactions account not only for the depletion of potassium or


sodium, but also for a decrease in pH because of the release of hydrogen
ions. The decrease in pH enables the water to dissolve metaIlic metals, to
convert bicarbonate to carbon dioxide, or to convert bisulfide to sulfide. The
Smackover brines often contain relatively high concentrations of sulfide.
Several investigators have attempted to determine what mechanism is
responsible for the increased concentration of caIcium and depletion of mag-
nesium relative to sea water in many subsurface brines. Chave (1960) and
Van Engelhardt (1960) compared ocean water with subsurface brines con-
taining high concentrations of caIcium, and demonstrated that dolornitiza-
tion cannot account for aIl of the calcium in the brine solutions. Van
Engelhardt (1960) noted that even the formation of chlorite utilizing magne-
sium with exchange of sodium and caIcium does not account for aIl of the
soluble caIcium; however, exchange reactions with other cIays were not con-
sidered. Kramer (1963) assumed that caIcium was more abundant in ancient
oceans, but White (1965) found this relation to be untenable and suggested
that shale-membrane filtration accounts for increased concentrations of
caIcium in some brines. Additional data are needed before more definite
conclusions can be made. The amounts and ratios of caIcium and magnesium
vary from one formation water to another as weIl as within one formation at
different geographic areas. Mineral formation, exchange reactions, leaching,
and shale-mernbrane filtration aIl can alter the composition of the brine.
However, in a specific area, one type of reaction may predominate.

Membrane-concentrated brines

EssentiaIly the postulate that clays and shales act as membranes to filter
dissolved solids from waters results from the fact that synthetic membranes
are used to desalinate waters by reverse osmosis. Conceivably, compacted
clays and shales rnay perform as imperfect semipermeable membranes. Solu-
RESEARCH STUDIES 241

tions of salts of different concentrations separated by a semipermeable


membrane will cause water from the lower salt-concentration side to move
through the membrahe to the higher concentration side, producing a greater
pressure on the high-concentration side. The pressure differential is the
osmotic pressure of the system and can account for abnormal pressures
found in some reservoirs.
Reverse osmosis occurs when hydraulic pressure in excess of the osmotic
pressure is applied to the high-concentration side, which forces water
through the membrane to the low-concentration side. The system is not
100% effective and some dissolved solids move through the membrane
(Kimura and Souriragan, 1967).
Such a system requires rather high pressure differentials in nature to
produce the highly concentrated brines found in some formations. The
osmotic pressure could produce pressure differentials in formations, but the
pressure comes to equilibrium as the two solutions equilibrate. The reverse
osmosis system works only as long as the excess hydraulic pressure is
applied. In the absence of the excess hydraulic pressure, the system comes to
equilibrium.
Larson (1967) reported some desalination results for water with reverse
osmosis using cellulose-acetate membranes. With a brackish water containing
about 4,300 mg/l of dissolved solids, input pressure of 42 kg/crn? and
temperature of 15.9°C, the ion rejection rates were as high as 99.9%. The
rejection order based on the percent rejected was:

Assuming that this mechanism operates in a shale filtration system, the order
of ion concentration on the high brine concentration side would be the
same. The ion concentrations on the fresher water side would be the reverse
or:

Other investigators have obtained similar results. For example, Loeb and
Manjikian (1965) found a rejection order of 804 -2 > Mg+2 > Ca+2 > Na"
> HC03 - > CI- > N03 -. Michaels et al. (1965) found a rejection order of
Ca+2 > Li" > Na" > K+ for the pressure independent portion of salt
transport in cellulose acetate reverse osmosis desalination membranes. This
correlates with the size of the hydrated ion radii because calcium is the
largest and potassium the smallest. Further, this indicates that the pore size
of the membrane is a controlling factor.
The data of Larson (1967) showed that sulfate and carbonate scale
formed on the high-pressure side of the membrane and if not removed would
cause flow to decrease or stop. The pH on the output or fresh-water side of
the membrane decreased.
242 OR,lGTN OF OTT,FTELD WAT RS

R ussell (1933) considered several processes which could produce subsur-


face brines more concentrated than sea water. He concluded that evapora-
tion of the water by natural gas generally is not important, water evapora-
tion incoarse-grained rocks generally is not important, gravitational settling
of dissolved solids is not greatly important, rocks containing considerable
amounts of feldspars and other unstable mineral s take up large quantities of
hydration water, clays adsorb bases and later expel them into solution
causing concentration, and osmosis may occur through semipermeable mem-
branes.
DeSitter (1947) noted that oilfield waters are altered as a result of two
prominent diagenetic phases. During the first phase magnesium, calcium,
sulfate, and carbonate precipitate from the original sea water. During the
second phase the concentration of rnagnesium and calcium ions increases
along with the concentration of other dissolved solids. He reasoned that the
second phase occurred because of filtration through semipermeable shales.
The filtration results because of sediment compaction until a semiper-
meable membrane develops which allows water molecules to pass through
but retards salt ions. Thus, the more concentrated brines are found where
sediment compaction and water flow distance were the greatest. This usually
occurs in the deepest portion of a basin.
McKelvey et al. (1957) forced aqueous saline solutions through ion-
exchange resins and found that the effluent solutions contained less dis-
solved salts than the influent solutions. Effluents from cation-exchange
resins were found to contain Na/K ratios similar to those in the influent;
however, the Mg/Ca ratios were at first higher than in the influent but with
additional squeezing the ratio decreased to much lower values. They
postulated that similar reactions occur during the compaction of sediments
to change the concentrations of constituents dissolved in waters.
Pressures of 7 kg/cm2 to 105 kg/crrr' were applied to force sodium
chloride solutions through cation-exchange membranes. The results indicated
that the membranes desalted the saline solutions, producing a filtrate con-
taining less salt than the influent. This salt filtering effect was attributed to
the electrical properties of the membrane.
Milne et al. (1964) determined the filtering efficiencies of sodium chloride
solutions by bentonite membranes. The filtration efficiencies were 94% at
140 kg/cm2 and 88% at 703 kg/cm2 with 0.5N sodium chloride. Increased
salinity caused less efficient filtration because filtration efficiencies of 94%
for 0.5N sodium chloride and 66% for 4N sodium chloride at a pressure of
352 kg/cm? were obtained. A similar mechanism could operate in the sub-
surface to create concentrated brines.
Young and Low (1965) performed an experiment using natural rock and
demonstrated that osmotic flow of water through shale and siltstone occurs.
The osmotic pressures produced were less than theoretical and they were
attributed to microcracks in the natural rock which caused them to be less
effective than a perfect membrane.
RESEARCH STUDIES 243

Bredehoeft et al. (1963) developed a mathematical model to predict the


distribution of ions within a formation. They assumed that a hydrostatic
head differential operates between the margin and center of a geologic basin,
producing a water movement upward through confining low permeability
beds. lf these low permeability beds contain clay membranes to restrict the
passage of ions, the waters on the upflow, or more permeable, side become
more concentrated in dissolved solids. They theorized that this process
produced the concentrated brines found in the Illinois Basin, and that their
model added weight to the membrane theory of brine concentration. A
major drawback to the model is the tremendous pressures that are necessary
to produce a movement of water upward through confining low permeability
beds.
Graf et al. (1965) found that isotopic fractionation occurred when waters
passed through shale micropores in the Illinois, Michigan, Alberta, and Gulf
Coast Basins. Their study did not yield sufficient evidence to estimate the
total fraction of water movement in the basins subsequent to sediment com-
paction. The 0180 concentrations in brin es did not indicate a direct corre la-
tion with ancient oceans.
A study of the oD o
and 180 in formation waters indicated that the water
was predominantly meteoric, little exchange or fractionation had occurred
to alter the deuterium, but extensive exchange between the water and rock
had altered the oxygen (Clayton et al., 1966). They postulated that forma-
tion waters in the Gulf Coast Basin lost their original connate water because
of sediment compaction and flushing, and that the present water is meteoric
water which carne in through outcrops.
This study was good; however, basic studies concerning the fractionation
and exchange of isotopes between water, hydrocarbons, and rocks need to
be made. Results of such studies should enable more positive interpretations.
A simplistic model was derived to determine the amounts of fresh water
and sea water necessary to create the brine compositions now present in the
Illinois and Michigan Basins (Graf et al., 1966). The model assumes: (1)
perfect efficiency of shale ultrafilters; (2) complete bacterial reduction of
sulfate with replacement in solution of equivalent bicarbonate; (3) complete
removal of bicarbonate and equivalent sodium by shale ultrafiltration; and
(4) magnesium reaction with calcium carbonate to form dolomite. The
dolomitization reaction furnished more soluble calcium than is possible for
the Illinois Basin, so another calculation was made assuming complete loss of
magnesium to clay minerals with no return of calcium.
The calculations indicated that less fresh water passed through the rocks
of the Illinois Basin than those of the Michigan Basin. These data conflicted
somewhat with Clayton et al. (1966) in that they argued that the water
molecules now in the Illinois Basin originated as fresh water, while the data
of Graf et al. (1966) indicated that too few volumes of fresh water passed
through the Illinois Basin to alter the brine significantly.
A study of the hydrodynamics of the Illíno ís Basin indicated that in
244 ORIGIN OF OILFIELD WATERS

recent times, before pumpage, the differences in vertical head in the deep
aquifers wer~ iusufficicnt to cause upward flow through shale, resulting in
ultrafiltration (Bond, 1972). In fact the head differentials were barely suf-
ñcíeni tu enablc upward f'low through an open conduit.
Berry (1969) outlined the relative factors that influence membrnne filtra-
tion in geologic environments. The membrane properties of shales are caused
by the electrical properties of their clays and organic materials. Clays
predominantly are cation exchangers with singly charged SiO- and
1
AlOSi- /4 sites and minor anion exchanges with replaceable OH- ions.
Divalent cations are adsorbed in preference to monovalent cations and
sodium is hyperfiltratcd with respect to lithium and strontium with respect
to calcium, because of preferential adsorption of ions with ionic potentials
most similar to the ionic potential of the exchange site. The selectivity of
hyperfiltration for the halogens is Cl ;;;.Br > I > F because of their substitu-
tion for OIl in the clays. Thus, in waters concentrated by this process the
Ca/Na, Na/Li, Sr/Ca, Cl/Br, Br/I, and I/F ratios should increase. These ratio
increases have been found in some brine systems, but by no means in all
systems.
Billings et al. (1969) found five types of formation waters in the Western
Canada Sedimentary Basin and postulated the origin of two of the types.
One type of water was formed by selective membrane filtration which pro-
duced waters containing high concentrations of dissolved solids. A second
type was a mixture of membrane-concentrated formation water and bitterns
formed after the precipitation of halite but before the precipitation of
sylvite. They theorized that the alkalies were filtered selectively by clay-
shale membranes, producing a concentrated brine, and that the relative con-
centration pattern is Rb > K > Na > Li. This pattern is the reverse of what
occurs by ion exchange but is similar to the surface mobilities of cations
along clay surfaces.
A detailed study of the Western Canada Sedimentary Basin, including a
determination of the rock volume and pore volume (Hitchon, 1968), the
effect of topography upon the fluid flow (Hitchon, 1969a), and the effect of
geology upon the fluid flow (Hitchon, 1969b), strongly suggested that
thermal, electro-osmotic, and chemico-osmotic forces are operating within
the basin to affect the fluid energy gradients. Pressure differentials of about
98 kg/cm 2 along with salinity differences of 200,000 mg/I between forma-
tions in close proximity were found which suggest that chemico-osmotic
forces are occurring.
Hitchon and Friedman (1969) used chemical analyses and stable-isotope
analyses for hydrogen and oxygen for surface waters, shallow ground waters,
and deep ground waters in a study of the origin of formation waters in the
Western Canada Sedimentary Basin. They postulated that surface waters have
mixed with diagenetically altered sea water to form the formation waters.
Using mass balance data for the deuterium and dissolved solid contents of
the formation waters, they calculated not only how much fresh water is
CONCLUSIONS 245

present in the modified sea water but also observed how it redistributed the
dissolved solids to produce salinity variations.
They concluded that formation waters result from mixing of surface
waters with modified mar in e 01' nonmarine water in the subsurface rocks,
that exchange of oxygen isotopes between the water and rock caused differ-
ent water types in different basins, and that f'ormation waters that have
passed through shale ultrafilters are more depleted in deuterium.
A study of the Surat Basin showed that most of its hydrocarbon accumu-
lations are associated with quasi-stagnant waters. The salinities of these
quasi-stagnant waters were higher than were the salinities of the waters in the
more dynamic recharge areas. The investigators postulated that these high
salinity waters were formed by membrane filtration because of cross-
formational flow and also that the hydrocarbon accumulations in these
quasi-stagnant are as resulted from release of hydrocarbons mo bilized by a
moving water. The hydrocarbons were released because of the higher
salinities of the waters in the quasi-stagnant areas (Hitchon and Hays, 1971).
A study of waters in sedimentary rocks of Neogene age in the northern
Gulf of Mexico Basin was made by Jones (1969). The hydrologic conditions
currently found in these sediments are similar to conditions that previously
occurred in older sedimentary basins. Osmotic flow has a dominant influence
upon the hydrology of normalIy and abnormally pressured aquifer systems
in the northern Gulf Basin.
Jones (1969) found that many forces such as gravity, sediment diagenesis,
different water salinities, ionic and molecular diffusion, different electrical
potentials of sediments, thermal potentials, pressure, and osmotic membrane
filtration affect the hydrology in this basin.
Fowler (1970) found that salinity variations within the Frio sands in the
Chocolate Bayou field, Brazoria County, Texas, are the result of selective
concentration of ions by shales acting as membranes. In this field, pressures
seem to reflect the flow paths of the waters, and the greatest changes in
pressures are found across shaly sections. Analyses of water samples from
this field over a 28-year period indicate decreasing salinity with production
time caused by dilution of the original brines by waters squeezed from the
shales adjacent to the aquifers.
Chilingarian and Rieke (1969) reviewed the processes which can alter the
chemical composition of formation waters. They concluded that most of the
original water was sea water, and that the concentration process in many
cases results from compaction and membrane filtration rather than evapora-
tion. Their experimental results indicated that solutions squeezed out of
rocks during compaction progressively decrease in dissolved solids concen-
trations with increasing depth,

Conclusions

The origin of oilfield waters is related to many natural processes. Initially,


r ••

246 ORIGIN OF OILFIELD WATERS

meteoric water reacted with weathered rack, soil, and organíc matt r. The
excess watcrs that did not penetrate the rock or soil caused the rock and soil
to erode and channels formed through which the water could mOYA more
easily. Forces of gravity causcd the water to move from areas of high poten-
tial to areas of low potential, and as the waters moved, the concentrations of
dissolved solids in them increased. Some of these waters found their way to
lakes and the sea. As they entered the lakes or seas their movement slowed,
causing some of the suspended particles in them to deposito Mixing of the
waters with the more saline waters in the sea caused dissolved carbonate and
organic compounds to precipitate.
Evaporation of the sea and lake waters caused other compounds such as
sulfates to precipitate. The pH of the waters changed slightly because of
reactions with the atmosphere, the sediments, and other waters. Each pH
change caused precipitation of compounds or dissolution of new corn-
pounds.
Some of the waters became highly concentrated in dissolved solids in the
more shallow marine environments. Evaporites formed in these lagoons,
pans, and exposed supratidal sabkhas. Evaporites also formed in deep-water
basins when the salinity of the water at the bottom of the basin became
sufficiently high.
The sediments were buried as additional sediments were deposited on
them, and water surrounding the sediment particles also was buried. As the
depth of burial increased, the sediments compacted and some of the water
was squeezed out. Both the squeezed-out water and the remaining interstitial
water reacted with minerals in the sediments to change the composition of
the dissolved solids in the water and the composition of the sediments.
Mechanisms that cause the oilfield waters to differ in composition from
water originally deposited with the sediments include ion exchange, infil-
trating waters, sediment leaching, mineral formation, sulfate reduction, and
ultrafiltration through clay-shale membranes.

References

Al'tovskii, M.E., Kuznetsova, Z.I. and Shvets, V.M., 1961. Origin o( Oi/ and Oil Deposits
(English Transl. by Consultants Bureau). Plenum Press, New York, N.Y., 107 pp.
Anonymous, 1964. Chemistry of the oceans. Chem. Eng. News, 42:12A.
Atwater, G.I. and Miller, E.E., 1965. The effect of decrease in porosity with depth on
fu ture development of oil and gas reserves in South Louisiana. Presented at A nnual
Meel., Am. Assoc. Peto Geol., New Orleans, La., 1965 - Bul/. Am. Assoc. Peto Ceo/.,
49 :334.
Ault, W.U., 1959. Isotopic fractionation of sulfur in geochemical processes. In: P.H.
Abelson (Editor), Researches in Geochemistry, John Wiley and Sons, ew York, N.Y.,
pp.241-259.
Baas Becking, L.G.M., Kaplan, I.R. and Moore, D., 1960. Limits of the natural environ-
ment in terms of pH and oxidation-reduction potentials. J. Geol., 68 :243-284.
REFERENCES 247

Baker, E.G., 1960. A hypothesis concerning the accumulation of sediment hydrocarbons


to form crude oil. Geochim. Cosmochim. Acta, 19 :309-317.
Barnes, H.L. and Czamanske, G.K., 1967. Solubilities and transpor t of ore minerals. In:
H.L. Barnes (Editor), Geochemistr y of Hy drot.herrnal Ore Deposits. Hol t, Rinehart,
and Winston, New York, .Y., pp.334-38l.
Bentor, Y.K., 1969. On the evolution of subsurface brines in Israel. Chem. Ceol.,
4:83-110.
Berr y, F.A. F., 1969. Relative factors influencing mernbrane filtration effects in geologic
environments. Chem. Ceol., 4: 295-30l.
Billings, G .K., Hitchon, B. and Shaw, D.R., 1969. Geochemistry and origin of formation
waters in th e Western Canada Sedimentary Basin, 2. Alkali metals. Chem. Ceol.,
4:211-223.
Birch, F., 1954. The present state of geothermal investigations. Geoph ysics, 19:645--659.
Bond, D.C., 1972. Hydrodynamics in deep aquifers of the IIlinois Basin. JIl. State Ceo!.
Suru., Circ., No. 470, 72 pp.
Borchert, H. and Muir, R.O., 1964. Salt Deposits - The Orig in, Me tamorphism, and
Deformation of Euapo rite. D. Van Nostrand, London, 338 pp.
Braitsch, O. and Herrmann, A.G., 1963. Zur Geochemie des Broms in Saliniiren Sedi-
menten, 1. ExperimenteIle Bestimmung der Br-Verteilung in Natürlichen Salzsystemen.
Geochim. Cosmo chim. Acta, 27 :361-39l.
Braitsch, O. and Herrmann, A.G., 1964. Zur Geochemie des Broms in Salinaren Sedi-
menten, 11. Die Bildungstemperaturen Primarer Sylvin- und Carnallit-Gesteine.
Geochim. Cosmo chirn. Acta, 28:1081-1109.
Bredehoeft, J.D., Blyth, C.R., White, W.A. and Maxe y, G.B., 1963. Possible mechanism
for concentration of brines in subsurface formations. Bull. Am. Assoc. Peto Geol;
47 :257-269. .
Brod, 1.0., 1960. On principal ru les in the occurrence of oil and gas accumulations in the
world. l ni. Ceol. Reu., 2:922-1005.
Broecker, W.S. and Takahashi, T., 1966. Calcium carbonate precipitation on the Bahama
Banks. J. Ceophys. Res., 71 :1575-1602.
Brongersma-Sanders, M., 1971. Origin of major cyclicity of evaporites and bituminous
rocks: an actualistic model. Mar. Geol., 11 :123-144.
Bruderer, W., 1956. Les océans souterrains fossiles et le pétrole. Assoc. Fr. Tech nol. Pelo
Bul!.,120:535-556.
Burst, J.F., 1969. Diagenesis of Gulf Coast clayey sediments and its possible relation to
petroleum migration. BulI. Am. Assoc. Peto Ceo/., 53:73-93.
Bush, P.R., 1970. Chloride-rich brines from sabkha sediments and their role in ore forma-
tion. l nst. Min. Me iall. Trans., 79:B137-B144.
Butler, G.P., 1969. Modern evaporite deposition and geochemistry of coexisting brines,
the Sabkha Trucial Coast , Arabian Gulf. J. Sedimento Petrol., 39 :70-89.
Caraway , W.H. and Gates, G.L., 1959. Methods for determining water contents of oil-
bearing formations. U.S. Bur. Min. Rep. l nuest.; No.5451, 81 pp.
Carpelan, L.H., 1957. Hydrobiology of the Alviso salt ponds. Ecology , 38:375-390.
Carpenter, A.B. and Miller, J.C., 1969. Geochemistry of saline subsurface water, Saline
County (Missouri). Chem. Geol., 4:135-167.
Cartrnill, J.C. and Dickey, P.A., 1970. Flow of a disperse emulsion of crude oil in water
through porous media. Bull. Am. Assoc. Peto Geol.; 54:2438-2447.
Case, L.C., 1955. Origin and current usage of the term "connate water". Bull. Am. Assoc.
Pelo Ceo/., 39:1879-1882.
Chave, K.E., 1960. Evidence on history of sea water from chemistry of deeper subsurface
waters of ancient basins. Bull. Am. Assoc. Peto Geol., 44:357-370.
Chebotarev, I.I., 1964. Metamorphism of natural waters in the crust of weathering.
Geochim. Cosmo chirn. Acta, 8.22 ·48,137~1?O, 198-21 .

248 ORIGIN OF OILFIELD WATERS

Chilingarian, G. V. and Rioke, El, H. H. 19h9. Some chemical altera tions of subsurface
waters during diagenesis. Chem. Geol.; 4:235-252.
Clark, S.P. and Ringwood, A.E., 1964. Density disturbance and constitution of the
mantle. Rev. Geo ph.y s.; 2:35-88.
Clarke, F.W. and Washington, H.S., 1924. The composition of the earth's crust. U.S.
Geol. Surv. Pro]. Paper, No.127, pp.I-112.
Clayton, R.N., Friedman, 1., Graf, D.L., Mayeda, T.K., Meents, W.F. and Shimp, N.F.,
1966. The origin of saline f'orrnat.ion waters, 1. Isotopic composition. J. Geophys.
Res., 71 :3869-3881.
Cloke, P.L., 1966. The geochemical application of Eh-pH diagrams. J. Ceo/. Educ.,
14:140-148.
Collins, A.G., 1967. Geochemistry of some Tertiary and Cretaceous age oil-bearing forma-
tion waters. Enoiro n. Sci. Techno l, 1:725-730.
Collins, A.G. 1969a. Chemistry of some Anadarko Basin brines containing high concen-
trations of iodide. Chem. Geol., 4:169-187.
Collins, A.G., 1969b. Solubilities of some silicate minerals in saline waters. U.S. Off
Saline Water Res. Dev. Progr. Rep., No.472, 27 pp.
Collins, A.G., 1970. Geochemistry of some petroleum-associated waters from Louisiana.
U.S. Bur. Min. Rep. Invest., No.7326, 31 pp.
Collins, A.G., 1972. Géochernical Classification of Formation Waters [or Use in Hydrocar-
bon Exploration and Production. M.S. Thesis, University of Tu lsa , Tu lsa , Okla., 63 pp.
Collins, A.G., Bennett, J.H. and Manuel, O.K., 1971. Iodine and algae in sedimentary
. rocks associated with iodine-rich brines. Geol. Soc. Am. Bull., 82:2607-2610.
Dapples, E.C., 1959. The behavior of silica in diagenesis. In: Silica in Sediments (A
Symposium) - Soc. Econ. Paleontol. Mineral., Spec. Publ., No.7, pp.36-54.
Deffeyes, K.S. Lucia, F.J. and Weyl, P.K., 1964. Dolomitization: observations on the
island of Bonaire, Netherlands Antilles. Science, 143 :678---B79.
Degens, E.T. and Chilingar , G.V., 1967. Diagenesis of subsur íace waters. In: G. Larsen
and G.V. Chilingar (Editors), Diagenesis in Sediments. Elsevier, Amsterdam,
pp.77-502.
Degens, E.T., Hunt , J.M., Reuter, J.H. and Reed, W.E., 1964. Data on the distribution of
amino acids and oxygen isotopes in petroleum brin e waters of various geologic ages.
Sedimeniology , 3: 199-225.
DeSitter, L.Y., 1947. Diagenesis of oilfield brines. Bull. Am. Assoc. Peto Geol.,
31: 2030-2040.
Dickey , P.A., 1969. Increasing concentration of subsurface brine with depth. Chem.
Ceo/.,4:361-370.
Dingman, R.J. and Angino, E.E., 1969. Chemical composition of selected Kansas brines
as a aid to interpreting change in water chemistry with depth. Chem. Geol.,
4:325-339.
Dunham, K.C., 1970. Mineralization by deep formation water s - a review. Inst. Metali.
Trans.,79:B127-B136.
Eckhardt, F.J., 1958. Über Chlorite in Sedimenten. Geol. Jahrb., 75:437-474.
Egleson, G.D. and Querio, C.W., 1969. Variation in the composition of brine from the
Sylvania formation near Midland, Michigan. Enoiron. Sci. Techriol.; 3 :367-371.
Ellis, A.J., 1968. Natural hydrothermal systems and experimental hot water/rock interac-
tion: reactions with NaCI solutions and trace metal extraction. Geochim. Cosmo chim.
Acta, 32:1356-1363.
Emer y, K.O., 1960. The Sea Off Southern California: A Modern Habitat of Petroleum.
John Wiley and Sons, New York, N.Y., 366 pp.
Erne ry , K.O. and Rittenberg, S.C., 1952. Early diagenesis of California Basin sediments in
relation to origin of oil. B ull. A m. Assoc. Peto Geol.; 36 :735-806.
REFERENCES 249

Evans, G., Kinsman, O.O.J. and Shearman, O.J., 1963 A reconnaissance survey of the
environment of the recent carbonate sedimentation along Trucial Coast, Persian Gulf.
In: L.M.J.U. van Straaten (Editor), Developments in Sedime nt ology, 1. Deltaic and
Shal/ow Marine Deposits. Elsevier, Amsterdam, pp.129-135.
Fowler, Jr., W.A., 1970. Pressures, hydrocarbon accumulation, and salinities - Chocolate
Bay ou field, Brazoria County, Texas. J. Peto Technol., 22: 411-423.
Friedman, G.M. and Sanders, J.E. 1967. Origin and occurrence of dolostones. In: G.V.
Chilingar, H.J. Bissel and R.W. Fairbridge (Editors), Carbonate Rocks - Origin, Occur-
rence and Classification. American Elsevier, New York, N.Y., pp.267-348.
Friedman, G.M. Fabricand, B.P., Imbimbo, E.S., Brey, M.E. and Sanders, J.E., 1968.
Chemical changes in interstitial waters from continental shelf sediments. J. Sedimento
Pe/rol.,38:1313-1319.
Garrels, R.M. and Christ, C.L., 1965. Solutions, Minerals, and Equilibria. Harper and
Row, New York, N.Y., 450 pp.
Gevirtz, J. L. and Friedman, G .M. 1966. Deep-sea carbonate sediments of the Red Sea
and their implications on marine Iithification. J. Sedimento Petrol., 36:143-151.
Ginsburg, R.N. and Lowenstam, H.A., 1958. Influence of marine bottom communities on
the depositional environment of sediments. J. Geol., 66:310-318.
Grabau, A.W., 1920. Geology of Nonmetallic Mineral Deposits Other Than Silicate, 1.
Principies of Salt Dep osits. McGraw-Hill, New York, N.Y., 435 pp.
Graf, O.L., Friedman, 1. and Meents, W.F., 1965. The origin of saline formation waters,
11. Isotopic fractionation by shale micropore systems. IlI. State Geol. Surv. Circ.,
No.393, 32 pp.
Graf, O.L., Meents, W.F., Friedman, 1. and Shimp, N.F., 1966. The origin of saline
formation waters, lII. Calciurn chloride waters. 1/1. State Geol. Surv. Circo No. 397, 60
pp.
Grim, R.E., 1952. Clay Mineralogy. McGraw-Hill, New York, N.Y., 396 pp.
Ham, W.E. 1962. Classification of carbonate rocks. Am. Assoc. Peto Geol., Mem. 1,279
pp.
Hanshaw, B.B. and Hill, G.A., 1969. Geochemistry and hydrodynamics of the Paradox
Basin region, Utah, Colorado, and New Mexico. Chem. Geol., 4: 263-294.
Helgeson, H.C., 1964. Complexing and Hydrothermal Ore Deposition. MacMillan, New
York, N.Y., 128 pp.
Hem, J.O., 1970. Study and interpretation of the chemical characteristics of natural
water. U.S. Ceo/. Surv. Water Supply Paper, No. 1473, 363 pp.
Hiltabrand, R.R. 1970. Experimental Diagenesis o] Argillaceous Sedimento Ph.O. Disser-
tation, Louisiana State University, Baton Rouge, La., 152 pp., unpublished.
Hitchon, B., 1968. Rock volume and pore volume data for plains region of Western
0
Canada Sedimentary Basin between latitudes 49° and 60 N. Bul/. Am. Assoc. Peto
Ceo/., 52: 2318-2323.
Hitchon, B. 1969a. Fluid f'lo w in the Western Canada Sedimentary Basin, 1. Effect of
topography. Water Resour. Res., 5:186-195.
Hitchon, B., 1969b. Fluid f'Iow in the Western Canada Sedimentary Basin, 2. Effect of
geology. Water Resour. Res., 5:460-46l.
Hitchon, B. and Friedman, 1., 1969. Geochemistry and origin of formation waters in the
Western Cana da Sedimentary Basin, 1. Stable isotopes of hydrogen and oxygen.
Geochim. Cosmochim. Acta, 33:1321-1349.
Hitchon, B. and Hays, J., 1971. Hydrodynamics and hydrocarbon occurrences, Surat
Basin, Queens land, Australia. Water Resour. Res., 7:658-676.
Ho lser , W.T. 1963. Chemistry of brine inclusions in Permian salt from Hutchison,
Kansas. In: J.L. Rau (Editor), Symposium on Salto Northern Ohio Geological Society,
Cleveland, Ohio, pp.86-103.
Hunt, J.M., 1968. Huw gas and oil Forrn l1l1d migl'llto. Wo)'Jd DiJ, 163'1 iil.O~l nO.
250 ORIGI T OF OILFIELD WATERS

Hunt, J.M. and Jamieson, G.W., 1958. Oil and organic matter in source rocks of petro-
leum. In: L.G. Weeks (Editor), Habitat o] Oil. American Association of Petroleum
Geologists, Tulsa, Okla., pp. 735-746.
Illina, L.V., 1954. Bahamancalcareoussands.Bul/.Am. Assoc. Peto Ceo/., 38:1-95.
Jones, B.F., Vandenburgh, A.S., Truesdell, A.H., and Rettig, S.L., 1969. Iuterst.itial brines
in playa sediments. Chem. Ceo/., 4:253-262.
Jones, P.H., 1969. Hydrology of Neogene deposits in the northern Gulf of Mexico Basin.
La. Water Resour. Res. Inst. Bul/., GT-2, 105 pp.
Kelley, W.P., 1948. Cation Exchange in Soils. Reinhold, New York, N.Y., 144 pp.
Khitarov, .1. and Pugin, V.A., 1966. Behavior of montmorillonite under elevated
temperatures and pressures. Ceochem. Int., 3 :621-626.
Kimura, S. and Souriragan, S., 1967. Analysis of data in reverse osmosis with porous
cellulose acetate membranes used. Am. Inst. Chem. Eng., 13:497-503.
Kozin, A.N., 1960. Geochemistry of bromine and iodine of formation waters in the
Kuybyshev area on the Volga. Peto Geol., 4:110-113.
Kramer, J.R., 1963. History of the composition of sea water - liquid inclusions com-
pared with a chemical equilibrium model. Ceo/. Soco Am., Spec. Paper, No. 73, 190 pp.
Kramer, J.R., 1969. Subsurface brines and mineral equilibria. Chem. Geol., 4 :37-50.
Krauskopf, K.B., 1956. Factors controlling the concentrations of thirteen rare metals in
sea water. Geochim. Cosmo chim. Acta, 9:1-32.
Krumbein, W.C., 1951. Occurrence and lithologic associations of evaporites in the United
States. J. Sediment Petro/., 21 :63-81. .
Krejci-Graf, K., 1963. Über Rurnanische Olfeldwasser. Ceo/. Mit t. Hydrogeo/. Hydrochem.,
2:351-392.
Kvenvolden, K.A., 1964. Hydrocarbons in modern sediments and the origin of petroleum.
Min, Mag., Colo. Schoo/ Min., 54:24-25.
Larson, T.J., 1967. Purification of subsurface waters by reverse osmosis. J. Am. Water
Works Assoc., 59:1527-1548.
Levorsen, A.r., 1966. Geology of Petro/euin. W.H. Freeman, San Francisco, Calif., 2nd
ed., 724 pp.
Loeb, S. and Manjikian, S., 1965. Six-rnon th field test of a reverse osmosis desalination
membrane. Ind. Eng. Chem. Process Design Dev., 4:207-212.
Lotze, F., 1938. Wichtigen Lager Stattender "nich t-erze ", IlI. Steinsalz und Kalisalze
Ceologie. Borntraeger, Berlin, 936 pp.
Mandl, 1., 1953. Solubilization of insoluble matter in nature, 11. The part played by salts
of organic and inorganic acids occurring in nature. Biochim. Biophys. Acta,
10:540-569.
Mandl, 1., Grauer, A. and Neuberg, C., 1952. Solubilization of insoluble matter in nature,
1. The part played by salts of adenosinetriphosphate. Biochim. Biophys. Acta,
8:654-663.
Manheim, F. T. and Bischoff, J.L., 1969. Geochemistry of pore waters on the Continental
Slope of the northern Gulf of Mexico. Chem. Ceol., 4:63-82.
McAuliffe, C., 1969. Determination of dissolved hydrocarbons in subsurface brines.
Chem. Geol., 4:225-234.
McKelvey, J.G., Spiegler, K.S. and Wyllie, M.R.J., 1957. Salt filtering by ion-exchange
grains and membranes. J. Ph.ys. Chem., 61(2):174-178.
Meyerhoff, A.A., 1970. Development in Mainland China, 1949-1968. Bull. Am. Assoc.
Pelo Ceol., 54:1567-1580.
Michaels, A.S., Bixler, H.J. and Hodges, Jr., R.M., 1965. Kinetics of water and salt
transport in cellulose acetate reverse osmosis desalination membranes. J. Colloid Sci.,
20:1034-1056.
Milne, I.H., McKelvey, J.G. and T'rump, R.P., 1964. Semi-permeability of bentonite me m-
branes to brines. Bul/. Am. Assoc. Peto Geol., 48:103-105.
REFERENCES 251

Moore, C.A~1969. The occurrence of oil in sedimentary basins. World Oil, 168:69-72.
Morris, R.C. and Dickey, P.A., 1957. Modern evaporite deposition in Peru. Bull. Am.
Assoc. Peto Geol., 41:2467-2474.
Neruchev, S.G. and Kovacheva, 1.S., 1965. The effect of geological conditions on the
amount of oil given up by source rocks. Dokl. Akad. Nauk U.S.S.R., 162:913-914.
Neumann, H.J. and Jobelius, H., 1967. Detection of emulsifying agents in crude oils and
oilfield waters as a contribution to the problem of oil migration. Erdol Kohle
Petrochem.,20:622-625.
Parker, J. W., 1969. Water history of Cretaceous aqui fers, East Texas Basin. Chem. Geol.,
4:111-133.
Peake, E. and Hodgson, G.W., 1966. Alkanes in aqueous systems, 1. Exploratory investiga-
tions on the accommodation of C2 o-C)) n-alkanes in distilled water and occurrence
in natural water systems. J. Am. Oil Chem. Soc., 43:215-222.
Pettijohn, F.J., 1957. Sedimentary Rocks. Harper and Brothers, New York, N.Y., 2nd
ed., 718 pp.
Philipp, W., Drong, H.J., Füchtbauer, H., Haddenhorst, H.G. and Jankowsky, W.J., 1963.
The history of migration in the Gifhorn Trough (NW Germany), Sixth World Peto
Congr., Frankfurt/Main, June, 1963, Sect. 1, Paper, No. 19, pp. 457-48l.
Philippi, G.T., 1965. On the depth, time and mechanism of petroleum generation.
Geochim. Cosmochim. Acta, 29:1021-1049.
Phleger, F.B. and Ewing, G.C., 1962. Sedimentology and oceanography of eoastal lagoons
in Baja, California, Mexieo. Geol. Soco Am. Bull., 73:145-18l.
Pirson, S.J., 1968. Redox log interprets reservo ir potential. Oil Gas J., 66 :69-75.
Plumley, W.J., Risley, G.A., Graves, Jr., R.W. and Kaley, M.E., 1962. Energy index for
limestone interpretation and classifieation. In: W.E. Hem (Editor), Classification of
Carbonate Rocks - Am. Assoc. Peto Geol., Mem.1, pp.85-107.
Pollard, T.A. and Reiehertz, P.O., 1952. Core-analysis praetiees - basie methods and new
developments. Bull. Am. Assoc. Peto Geol., 36:230-252.
Quaide, W., 1958. Clay minerals from salt eoneentration ponds. Am. J. Sci., 256 :431-437.
Rankama, K. and Sahama, T.G., 1950. Geochemistry. Chicago University Press, Chicago,
Ill., 991 pp.
Riley, G.A., 1944. The earbon metabolism and photosynthetie efficieney of the earth as a
whole. J. Am. Sci., 32 :134.
Rittenhouse, G., 1967. Bromine in oilfield waters and its use in determining possibilities
of origin of these waters. Bull. Am. Assoc. Peto Geol., 51: 2430-2440.
Rittenhouse, G., Fulton, R.B., Grabowski, R.J. and Bernard, J.L., 1969. Minor elements
in oilfield waters. Chem. Geol., 4:189-209.
Rosenqvist, 1.T., 1962. The influence of physico-ehemical factors upon the meehanical
properties of clays. Clays Clay Minerals, 9: 12-27.
Ross, C.S., 1943. Clays and soils in relation to geologic processes. J. Wash. Acad. Sci.,
33: 225-235.
Russell, W.L., 1933. Subsurfaee eoncentration of ehloride brines. Bull. Am. Assoc. Peto
Geol.,17:1213-1228.
Schoeller, H., 1955. Geoehemie des eaux souterraines. Rev. 1nst. Fr. Pet., 10 :181-213,
219-246, 507-552.
Siever, R., Beck, K.C. and Berner, R.A. 1965. Composition of interstitial waters of
modern sediments. J. Geol., 73:39-73.
Skinner, B.J., 1969. Earth Resources. Prentiee-Hall, Englewood Cliffs, N.J., 149 pp.
Sloss, L.L., 1953. The signifieanee of evaporites. J. Sedimento Petrol., 23:143-16l.
Smith, P.V., 1954. Studies on oriain of petroleum: or-cnrre nce of hyrlroearbons in reeent
sediments. Bull. Am. Assoc. Peto Geol., 38:377-38l.
Stumm, W. and Morgan, J.J., 1970. Aquatic Chemistry . Wiley-Interseience, Div. of John
Wiley and Sons, New York, N.Y., 583 pp.
252 ORTGIN OF OILFIELD WATERS

Taylor, D.W., 1956. Jt'undam nrurs tJ( Soil Mcchanice, John Wi)py an d Sons, New York,
N.Y., 700 pp.
Terzaghi, K. and Peck, R. B., 1968. Soil Mechonics in Engineering Practice. John Wiley
and Sons, New York, .Y., 84 pp.
T'oorns, J.S., 1970. Revicw o f knowledge of metalliferous brines an d related deposits.
Inst. Min. Metal/. Trans., 79:B116-B126.
Truesdall, A.H. and Jones, B.F., 1969. Ion association in natural brines. Chem. Ceo/.,
4:51-62.
Usiglio, J., 1849. Analyse de l'eau de la Mediterranée sur les cates de France. Ann. Chim.
Phys., 3:92-107, 27 :172-191.
Valyashko, M.G., 1956. The geochemistry of bromine in halogenesis processes and the
use of bromine con tent as a gene tic and prospecting criterion. Geochemistry ,
1 :33-49.
Van Everdingen, R.O., 1968. Studies on formation waters in western Canada; geo-
chemistry and hydrodynamics. Can. J. Earth Sci., 5: 523-543.
Van Nostrand Press, 1958. Science Encyclopedia. Princeton, N.J., 2nd ed., pp. 371-1528.
Vinogradov, A.P., 1953. The elementary chemical composition of marine organisms.
Sears. Found. Mar. Res., Mem., 11:85,91,107,216.
Visher, G.S., 1965. Use of vertical profile in environmental reconstruction, Bul/. Am.
Assoc. Peto Geol., 49:41-61.
Von Engelhardt, W., 1960. On the chemistry of the pore solution of sediments. Uppsa/a
Univ. Ge ol. Inst. Bul/., 40: 189-204 (in German).
WelJer, J.M., 1959. Compaction of sediments. Bul/. Am. Assoc. Peto Geol., 43:273-310.
White, D.E., 1957. Magmatic, connate, and metamorphic waters. Geol. Soc. Am. Bul/.,
68:1659-1682.
White, D.E., 1965. Saline waters of sedimentary rocks. In: A. Young and J.E. GalJey
(Editors), Fluids in Subsurftice Environmenls - Am. Assoc. Peto Geol., Mem.4,
pp.342-366.
Young, A. and Low, P.F., 1965. Osmosis in argilJaceous rocks. Bui/. Am. Assoc. Peto
Ceo!., 49:1004-1008.
Chapter 8. CLASSIFICATION OF OILFIELD WATERS

Classification of waters provides a basis for grouping closely related


waters. Because the grouping is chemical, it is dependent upon the dissolved
constituents found in the waters. Most of the classification systems devel-
oped to date have considered only the dissolved majar inorganic constituents
and have ignored the organic and the minar and trace inorganic constituents.
Waters as related to the earth are meteoric, surface, and subsurface. Sur-
face waters can be fresh or saline if the amounts of dissolved constituents in
the waters are used to classify them. For example, water from melting snow
on a mountain top usually will contain small amounts of dissolved mineral
matter and can be classified as fresh water, while water in an ocean will
contain about 35,000 mg/l dissolved minerals and is classified as saline.
Waters found in rivers connecting the mountain stream to the ocean may
contain varying amounts of dissolved constituents and depending upon the
amounts can be classified a,~fresh or saline. In a similar manner, subsurface
waters are classified as fresh or saline. Merely classifying a water as either
fresh or saline does not provide a very useful classification. The dissolved
constituents that are used in many classification systems depend upon the
amounts or ratios of sodium, magnesium, calcium, carbonate, bicarbonate,
sulfate, and chloride found in the water. The reason for this is that these are
the ions that usually are determined or calculated in a water. (Sodium often
is calculated from the difference found in the stoichiometric balance of the
determined anions and cations.)
The amounts and ratios of these constituents in subsurface waters are
dependent upon the origin of the water and what has occurred to the water
since entering the subsurface environment. For example, some subsurface
water s found in deep sediments were trapped during sedimentation, while
other subsurface waters have been diluted by infiltration of surface waters
through outcrops. Some waters have been replaced by infiltration water.
Also, rocks containing the waters often contain soluble constituents, which
dissolve in the water s or contain chemicals which will exchange with chemi-
cals dissolved in the waters causing alterations of the dissolved constituents.
The amounts of dissolved constituents found in subsurface waters can
range from a few milligrams per liter to more than 350,000 mg/l. This
salinity distribution is dependent upon several factors, including hydraulic
gradients, depth of occurrence, distance from outcrops, mobility of the
dissolved chemical elements, soluble material in the associated rocks, and the
exchange reactions.
254 CLASSIFICATION OF OILFIELD WATERS

Portions of three classification systems (Palmer, 1911; Sulin, 1946;


Schoeller, 1955) and Bojarski's (1970) modification of Sulin's system were
applied to about 4,000 formation waters (U.S. Bureau of Mines, 1965). The
waters were analyzed by standard methods (American Petroleum Institute,
1968). The results indicated that the classifications are useful in exploration
and production problems.

Palmer's classification

Palmer (1911) observed that the basic characteristics of natural waters are
dependent upon their salinity (salts of strong acids) and alkalinity (salts of
weak acids). Salts that cause salinity are those that are not hydrolyzed, while
alkalinity is caused by free alkaline bases produced by the easily hydrolyz-
able salts of weak bases.
AII positive ions (cations) including hydrogen can cause salinity, but of
the negative ions (anions), only the strong acids, (e.g., chloride, sulfate, and
nitrate) can cause salinity. Because salinity is dependent upon the combined
activity of the cations and anions and is limited by the reacting values of the
strong acids, its value is determined by multiplying the total value of the
strong acids by two.
Alkalinity is caused by free alkaline bases as a result of the hydrolytic
action of water on dissolved bicarbonates and other weak acid salts. The
alkalinity value is calculated by doubling the reacting values of the bases
which exceed the reacting values of the strong acids.
The ions that commonly are found in waters comprise three groups: (a)
alkalies (sodium, potassium, lithium), whose salts are easily soluble in water
and do not cause hardness; (b) alkaline earths (magnesium, calcium, stron-
tium, barium), whose salts cause hardness and many of which are sparingly
soluble; and (e) hydrogen, whose salts are acids and cause acidity.
Geologists know what "strong alkalies", "alkaline earths", "strong acid
radicles", "weak acid radicles", "ions", and "reacting values" mean general-
Iy. To compare several analyses it usually is easier if they are made on a
chemical basis. The, proportions of the various ions do not react in propor-
tion to the various weights given in milligrams per liter but rather in propor-
tion to their "capacity for reaction", or "reaction value". The reacting value
of each ion is determined by multiplying the amount of each radicle by
weight (mg/I) by its "reaction coefficient", which is the valence of a radicle
divided by its atomic weight.
The groups of the ion s are determined by summing the reacting values of
their members, and according to the predominance of reacting values of the
groups, five special properties were designated by Palmer. To determine the
special properties, the reacting values of a group of cations or anions are
doubled so that the full value of a given special property is considered. The
terms "primary" and "secondary" were used to qualify the general proper-
PALMER'S CLASSIFICATION 255

ties of the water; e.g., the principal soluble decomposition products of the
oldest rock formations are the alkalies (primary), while more recent rock
formations are the principal source of the alkaline earths (secondary). This
theory of Palmer's that the terms primary and secondary are associated with
the age of the rock should not necessarily be considered undisputably true,
beca use primary salinity certainly can be acquired from other soluble
material than that derived directly from decomposition products of the
oldest rock formations.
The five special properties of water are:
(1) Primary salinity (alkali salinity); that is, salinity not to exceed twice
the sum of the reacting values of the radicles of the alkalies.
(2) Secondary salinity (permanent hardness); that is, the excess (if any) of
salinity over primary salinity, not to exceed twice the sum of the reacting
values of the radicles of the alkaline earths group.
(3) Tertiary salinity (acidity); that is, the excess (if any) of salinity over
primary and secondary salinity.
(4) Primary alkalinity (perrnanent alkalinity); that is, the excess (if any) of
twice the sum of the reacting values of the alkalies over salinity.
(5) Secondary alkalinity (temporary alkalinity); that is, the excess (if any)
of twice the sum of the reacting values of the radicles of the alkaline earths
group over secondary salinity.
Reacting values in percent are used in this system. The percentage values
are determined by summing the milliequivalents of all the ions, dividing the
milliequivalents of a given ion by the sum of the total milliequivalents, and
multiplying by 100. Waters are classified by numerical values of the relation-
ships of anions to the cations, where a, b, and d represent the percentage
values of the alkali cations, alkaline earth cations, and strong acid anions,
respectively. Any one of the following five conditions may exist: d may be
equal to or less than a, greater than a and less than a + b, equal to a + b, or
greater than a + b. Using these conditions, waters are classified into five
classes:

Class 1: d < a
2d = primary salinity
2(a - d) = primary alkalinity
2b = secondary alkalinity

Class 2: d =a
2a or 2d = primary salinity
2b = secondary alkalinity

Class 3: d > a; d < (a + b)


2a = primary salinity
2( d - a) = secondary salinity
2(a + b - d) = secondary alkalinity
256 CLASSIFICATION OF OILFIELD WATERS

Class 4: d = (a + b)
2a = primary salinity
2b = secondary salinity

Class 5: d > (a + b)
2a = primary salinity
2b = secondary salinity
2( d - a - b) = tertiary salinity (acidity)

These five classes of water are found in nature. Examples of the first three
classes are various surface waters, sea water and brines represent class 4,
while mine drainage waters and waters of volcanic origin fall in class 5
(Palmer, 1911).
Rogers (1917, 1919) studied oilfield waters of the San Joaquin Valley,
California, and used the classification system of Palmer (1911). He found
that generally the surface waters of the San Joaquin Valley possess second-
ary salinity rather than primary alkalinity, contain more sulfate than
chloride, and contain low amounts of bicarbonate. With increasing depth,
the subsurface waters decrease in secondary salinity until primary alkalinity
becomes evident. Waters above an oil zone often contained hydrogen sulfide,
which was attributed to reduction of sulfates by hydrocarbons, thus de-
creasing the amounts of sulfate and increasing the bicarbonate in the water,
which Rogers called an altered water. Further he found that, in these altered
water s in close proximity to hydracarbon accumulations, chloride becomes
relatively and absolutely important beca use of the residual chloride from the
original (ancient) sea water chlorides as compared to waters above the oil
zone which often are freshened because of a more hydrodynamic situation.
Altered waters, according to his definition, can have either primary alkalinity
or secondary salinity depending upon their amounts of carbonate and chlo-
ride, but normal waters have only secondary salinity.
Elliott (1953) used the Palmer system to determine the chemical charac-
teristics of some Paleozoic age formation waters in Texas. He found that all
of the waters in the group that he studied (about 70) contained predomi-
nant, primary salinity. Many of these water s contained appreciable concen-
trations of sulfate; one contained 5,800 mg/l sulfate, and many contained
more than 2,000 mg/I. The calcium concentration ranged up to 13,000 mg/l
while the bicarbonate concentrations ranged up to 800 mg/l.
Ostroff (1967) used the Palmer classification to classify waters fram
several basins and to compare this classification system with two other
systems. He found that the Palmer system groups some of the constituents
together that are not closely related chemically. Furthermore the system
does not consider ionic concentrations or saturation conditions related to
sulfate or bicarbonate.
SULIN'S CLASSIFICATION 257

Sulin 'S classification

Sulin (1946), a Russian geochemist, proposed a classification system based


upon various combinations of dissolved salts in the waters. The waters are
described according to chemical type, subdivided into group, subgroup, and
class. He found four basic environments of natural water distribution:
(1) Continental (terrestrial) conditions which pro mote the formation of
sulfate waters. Such conditions supply soluble sulfate constituents to the
water and the genetic type of such a water is "sulfate-sodium ".
(2) Continental conditions which pro mote the formation of sodium bicar-
bonate waters. The genetic type is "bicarbonate-sodium ",
(3) Marine conditions and the formation of a "chloride-magnesiurn " type
of water.
(4) Deep subsurface conditions within the earth's crust and the formation
of a "chloride-calciurn" type of water.
The first two types are characteristic of meteoric andjor artesian waters,
the third of marine environments and evaporite sequences, and the fourth of
deep stagnant conditions.

Ty pes, groups, and subgroups

Water composition is expressed in milligram-equivalents of the separable


ions, and the composition is calculated per 100 g of water. The percent of
the sum of the equivalents is used to exclude the degree of water mineraliza-
tion, and to compare waters containing different amounts of dissolved solids.
The ratio NajCl expressed in the percent equivalent form determines the
genetic water type. If the value is greater than one, sodium predominates
over chloride and the excess sodium can be combined with sulfate or bicar-
bonate. Therefore waters with a NajCl ratio greater than one belong to the
bicarbonate-sodium or the sulfate-sodium types. Sulin calculated sodium as
the sum of all the alkalies (Li, K, Na etc.) and chloride as the sum of all the
halides (Cl, Br, 1).
The ratio (Na - Cl)jS04, if greater than one, indicates that the water is
the bicarbonate-sodium type, while if it is less than one it is the sulfate-
sodium type. Similarly the ratio (Cl- Na)jNa if less than one indicates the
chloride-magnesium type, but if greater than one it indicates the chloride-
calcium type.

Water classes

Subdivision of the groups of waters were made by Sulin (1946) using the
Palmer (1911) characteristics, because these characteristics express the dis-
solved constituents in the waters in a generalized formato For example, the
sum of the alkali chlorides and sulfates corresponds to primary salinity, and
the sum of the alkaline earth chlorides and sulfates corresponds to secondary
258 CLASSIFICATION OF OILFIELD WATERS

salinity and the sodium bicarbonate-calcium stage. No sodium bicarbonate is


present in sulfate-sodium, chloride-magnesium, or chloride-calcium types of
water; therefore, these types are classified as follows:
(1) Class A2 : secondary alkalinity predominates (alkaline earth carbonates
and bicarbonates).
(2) Class S2: secondary salinity predominates (alkaline earth sulfates and
chlorides) .
(3) Class SI: primary salinity predominates (alkali sulfates and chlorides).
(4) Class S3: tertiary salinity predominates (iron and aluminum sulfates
and chlorides and free strong acids).
Bicarbonate-sodium type waters contain sodium bicarbonate and are
classified as follows:
(5) Class A2 : secondary alkalinity predominates (alkaline earth carbonates
and bicarbonates).
(6) Class A¡ : primary alkalinity predominates (alkali carbonates and bicar-
bonates).
(7) Class SI: primary salinity predominates (alkali chlorides and sulfates).
(8) Class A3 : tertiary alkalinity predominates (iron and aluminum carbon-
ates and bicarbonates).
The water classification is expressed by use of a formula representing
decreasing values of the Palmer characteristics. For example, S¡ S2 A2 indi-
cates that primary salinity is predominant and is followed by secondary
salinity and secondary alkalinity. Therefore, the classes are subdivided into
subclasses, and class S 1 can include the subclass SI S2 A2, SI A2 82, SI S2'
and S¡. Table 8.1 outlines Sulin's method of water characterization. Table
8.II briefly outlines the relative values of the coefficients which determine
the four genetic types of waters.
The Palmer characteristics do not account for the interrelations between
chloride and sulfate and between calcium and magnesium. Therefore, Sulin
calculated the ratio S04 jCl and Ca/Mg to establish additional subgroups. The
complete water characterization included the following: (a) water formula
given in Palmer characteristics; (b) coefficients in percent equivalents for
SO4jCl and Ca/Mg; (c) sum of the milligram equivalents per 100 g of water
(L r) to illustrate the degree of water mineralization; and (d) the genetic
coefficients (Na - Cl)j S04 and (Ca - Na)jMg to determine the water type,
and NajCl to determine related genetic types of water.

Hydrochemical indicators of hydrocarbons

8ulin (1946) noted that certain properties of subsurface water s were


favorable indicators of hydrocarbon accumulations. The bicarbonate-sodium
and chloride-calcium types of waters are widely found in oilfields. However,
the chloride-calcium type is the more favorable indicator if it has the most
characteristic composition plus certain minor or micro constituents. In gen-
eral, he determined that hydrocarbon accumulations are most commonly
8ULIN'8 CLA88IFICATION . 259

TABLE 8.1

8ulin's method of water characterization

NafCl > 1

Sultate-sodium
l'
ty pe: (Na + -
804
Cl-)
2
<1 Bicarbonate-sodium type:
(Na+-cn>
80 -2 1
4

Bicarbonate group Bicarbonate group


class A2 class Al
calcium subgroup sodium subgroup
magnesium subgroup class A2
8ulfate group calcium subgroup
class 81 magnesium subgroup
calcium subgroup sodium subgroup
magnesium subgroup class 81
sodium subgroup sodium subgroup
class 82 8ulfate group
calcium subgroup class 81
magnesium subgro up sodium subgroup
Chloride group Chloride group
class 81 class 81
calcium subgroup sodium subgroup
magnesium subgroup
sodium subgroup

NafCl < 1

(Cl- - Na+) (Cl- - Na+) >


Chloride-magnesiurn type: Mg+2 <1 Chloride-calcium type: Mg+2 1

Bicarbonate group Bicarbonate group


class A2 class A2
calcium subgroup calcium subgroup
magnesium subgroup magnesium subgroup
8ulfate group 8ulfate group
class 81 class 81
calcium subgroup calcium subgroup
magnesium subgroup magnesium subgroup
class 82 class 82
calcium subgroup calcium subgroup
magnesium subgroup magnesium subgroup
Chloride group Chloride group
class 81 class 81
calcium subgroup calcium subgroup
magnesium subgroup magnesium subgroup
sodium subgroup sodium subgroup
class 82 class 82
calcium subgroup calcium subgroup
magnesiurn subgroup magnesium subgroup
260 CLASSIFICATION OF OILFIELD WATERS

TABLE s.n
Coefficients characterizing the genetic types of waters

Type of water

Chloride-calcium <1 <O >1


Chloride-magnesium <1 <O <1
Bicarbonate-sodium >1 >1 <O
Sulfate-sodium >1 <1 <O

related to water types in this order: chloride-calcium > bicarbonate-sodium


> chloride-rnagnesium > sulfate-sodium. Most oilfield waters of the
chloride-calcium type belong to the S] S2 A2 class with a few in the S2 SI
A2 class, while most oilfield bicarbonate-sodium waters belong to the S] A]
A2 and A] S] A2 classes.
Other significant indicators were grouped by Sulin; however, none of
them can assure the existence of a hydrocarbon deposit, and certainly they
cannot pro vide definite evidence of the size of the accumulation. The groups
are as follows:
Group 1: direct hydrocarbon indicators; for example, naphthenic acid salts
and iodide. The naphthenic acids are more soluble in bicarbonate-sodium
type waters and are related to the composition of the hydrocarbon accumu-
lation. Iodide is related to oi! because it must have an organic origino Sulin
also noted the dissolved gases in the waters and considered the heavier
hydrocarbons such as ethane and butane and the absence of oxygen as direct
indicators.
Group II: highly mineralized chloride-calcium or bicarbonate-sodium
types of water containingreduced forms of sulfur are important indirect
indicators of oil. The sulfate content should be low to indicate interaction
with bituminous constituents and/or sulfate-reducing bacteria.
Group III: in this group are constituents which have no genetic relation-
ship to hydrocarbons but appear characteristic of waters that are related to
hydrocarbon accumulations. The constituents are bromide, boron, barium,
strontium, radium, and possibly fluoride.

Modification of Sulin's system by Bojarski

Bojarski (1970) studied 400 water analyses and differentiated hydro-


chemical zones within basins in Poland that appear suitable for preservation
of hydrocarbon deposits. He distinguished the waters as follows:
(1) Waters of the bicarbonate-sodium type. Such waters occur in the
upper zone of a sedimentation basin, with "intense water exchange" (that is,
a hydrodynamic situation where the waters are moving at a relatively fast
geological rate), which promotes unfavorable conditions for the preservation
of petroleum and natural gas deposits. The waters are defined by the ratio
MODIFICATION OF SULIN'S SYSTEM 261

(Na - Cl)jS04 > 1. As Sulin (1946) noted, if the ratio Na/Cl in epm is
greater than 1, the water contains more sodium than chloride and the excess
sodium can react with sulfate or bicarbonate ions. Therefore, such waters
belong to the bicarbonate-sodium or sulfate-sodium types. If the ratio
(Na - Cl)jS04 is greater than 1, it indicates an excess of sodium with respect
to both chloride and sulfate.
(2) Waters of the sulfate-sodium type with (Na - Cl)jS04 < 1. This ratio,
if less than 1, indicates that all of the sodium will react with chloride or
sulfate.
(3) Waters of the chloride-magnesium type with (Cl - Na)jMg < 1. A ratio
of this type indicates that all of the chloride will react with sodium and
magnesiun. Such a water is characteristic of the transition zone between a
hydrodynamic area which is becoming more hydrostatic in the deeper part
of the basin, and the amount of dissolved bromide increases directly with the
(Cl- Na)jMg ratio.
(4) Waters of the chloride-calcium type with (Cl - Na)jMg > 1. This ratio
indicates an excess of chloride with respect to sodium and magnesium, and
the excess will react with calcium. This type of water occurs in deeper zones
which are isolated from the influence of infiltration waters and are hydro-
static or almost hydrostatic.
Bojarski observed a large variation in the chemical composition in the
chloride-calcium type of water and subdivided this type as follows:
(a) The first class, chloride-calcium I with Na/Cl > 0.85 characterizes an
active hydrodynamic zone with considerable water movement. It is con-
sidered a zone of little prospect for the preservation of hydrocarbon
deposits.
(b) The second class, chloride-calcium 11 with Na/Cl = 0.85-0.75, charac-
terizes the transition zone between an active hydrodynamic zone and a more
stable hydrostatic zone of the sedimentation basin, which is generally con-
sidered a poor zone for hydrocarbon preservation.
(e) The third class, chloride-calcium III with Na/Cl = 0.75-0.65 (0.60),
characterizes favorable conditions for the preservation of hydrocarbon
deposits.It is designated as a fairly favorable environment for the preserva-
tion of hydrocarbons.
(d) The fourth class, chloride-calcium IV with Na/Cl = 0.65-0.50, is
characterized by complete isolation of the hydrocarbon accumulations as
well as by the presence of residual waters. It is considered a good zone for
the preservation of hydrocarbons.
(e) The fifth class, chloride-calcium V with Na/Cl < 0.50, is characterized
by the presence of ancient residual sea water which has been highly altered
since original deposition, both in the concentration of dissolved solids and in
the ratios of the dissolved constituents. Bojarski considers a zone of this type
to be one of the most likely areas where hydrocarbons are accumulated.
Additional characteristics of water associated with hydrocarbon accumula-
tions are as follows: (1) iodide > 1 mg/I; (2) bromide > 300 mgjl (increasing
262 CLASSIFICATION OF OILFIELD WATERS

iodide and bromide concentrations may point to a bitumen accumulation);


(3) ratio CI/Br < 350; and (4) 804 x 100/CI < 1.
In addition to indicating the degree of alteration, bromide and iodide as
biophile constituents play a decisive role in the classification Bojarski
adopted. This followed because of the increased concentration of biophile
elements in the waters accompanying a petroleum deposito The concen-
tration of iodide in the ground waters depends mainly on the organic sub-
stances, whereas the concentration of bromide up to a certain limit takes
place in an inorganic medium, but an increase in bromide must be evaluated
as a positive indication. In many waters accompanying petroleum deposits,
large amounts of bromide and smaller amounts of iodide were detected, or
vice versa. This probably is related to the type of bituminous substances
which absorb the individual biophile elements in different amounts.

Chebotarev's classification

Chebotarev (1955), an Australian geochemist, classifies waters on the basis


of dissolved bicarbonate, sulfate, and chloride, and he does not consider the
acid waters or those that contain free sulfuric or hydrochloric acid. His
fundamental assumption is that the anions are independent variables while
the cations are dependent.
The geochemical types of waters are related to the products of weath-
ering. Table 8.lI1 illustrates the cycles and products that are produced by
weathering. During the first cycle the igneous rocks are weathered allowing
chloride, sulfate, calcium, sodium, silica, and magnesium to go into solution.
The second cycle is the weathering of sedimentary rocks with the solution of
more of the same products. The third cycle is the weathering of recent drift
and yields of the above constituents plus aluminum and iron.
Table 8.IV illustrates Chebotarev's (1955) geochemical classification of
subsurface waters. The phase of weathering corresponds to four phases of
the solution and redistribution of the chemical constituents in the earth 's
crust and correlates with their relative mobilities. He plotted the relative
mobilities of nine chemical constituents using the mobility of chloride as
100%. From this four phases were obtained, namely: (1) chloride and sulfate
100% to about 58% mobility; (2) calcium, sodium, magnesium and potas-
sium 3% to about 1.2% mobility; (3) silica about 0.20% mobility; and (4)
iron oxide and aluminum oxide, less than 0.05% mobility. The four phases
of weathering correspond to the products of weathering shown in Table 8.IV
and also to the cycles and products of weathering in Table 8.IlI. For
example the fourth phase in Table 8.IV corresponds with the first cycle in
Table 8.lII.
The genetic types of water shown in the upper portion (A) of Table 8.IV
do not correspond directly with the weathering phases since the genetic
types overlap the phases. These genetic types are related to the accumulation
products shown in Table 8.lII. In the lower portion (B) of Table 8.IV are the
(")
::r:
i?'1
t:O
O
>-3
>
::o
i?'1
<
TABLE S.III ri5
o
Cycles and types of products of weathering (after Chebotarev, Hi55)
r-
>
o:
Cycles of First cycle (orthoeluvium) Second cycle (paraeluvium) Third cycle (Neoluvium) co
weathering: from igneous and highly metamorphosed rocks from sedimentary rocks from Recent Drift
:;;
(=3
residual products accumulative products residual products accumulative products resid ual products accumulative products >
>-3
Types of (1) coarse detrital (1) chloride-sulphate (1 ) detri tal (1) chloride-sulphate (1) solonez and (1) chloride-sulphate (3
products (chiefly alluvial) gypsum bearing Z
of weathering (2) calcareous (2) calcareous (2) siallitic (2) calcareous (2) leached supra- (2) CaC03
(under vegetation (chiefly colluvial (supra·calcareous chloride-sulphate calcareous
cover) and proluvial) (siallitic)
(3) siallitic (3) siallitic (3) allitic (?) (3) unsaturated siallitic·2 (3) unsaturated siallitic (3) alumino· and ferri·
calcareous chloride- siallitic·2 siliceous system
sulphate

(4)allitic*3
(lateritic crust of
weathering)

.1 A large quantity of silica and much of its calcium and sodium compounds are temoved; the aluminosilicates pass gradually into residual aluminosilicic acids,
i.e., acids of the kaolin type .
• 2 The action of carbonated atmospheric water is insufficient to replace the absorbed ions by hydrogeri .
• 3 The accumulation of sesquioxides at the expense of the leaching out oí the alkalis, alkaline earth, and silica.

m
~
v.o
TABLE S.IV m
~
*""
Geoehemieal classif'ication of subsurfaee waters (after Chebotarev, 1955)

(A) Relotionship o] the products o{ weathering to the genetic t y pes o] water

Prcsumable phase of weathering Products of weathering Genetic types of water

Fourth phase residual (orthoeluvium and detrital paraeluvium) bicarbonate (alkaline)


bicarbona te-ehloride (alkaline-sali ne)
Third and partly seeond phases sia lIitie dri ft
ehloride-bicarbonate (sali ne-al kaline )
Second and part ly first phases calcareous accumulation
ehloride-sulfate (saline)
First phase ehloride-sulphate aeeumulation chloride (saline )

(B) Geochemical groups o] waters

Major group Class Genetie types of Reacting value in percent


of water water
-2
HCO) - + CO, - CI- S04 CI- + SO.-2 HCO) - + CI- HC03 - + S04 -2 (J
r-
Biearbonate bicarbonate >40 <10 Na+ + K+ prevail in all types of >
en
wat~rs
(',+ and Mg+2 less than 2.5 in the
g;
>-xJ
11 biearbo na te-eh loride 40-30 10-20 water of high saline facies and less
than 19.0 in low saline facies
(3
III eh loride-biearbona te 30-15 20-35 >
>-3
(3
Sulphate IV
Z
sulphate-ehloride 15-5 <25 >25 Na + + K+ prevail in all types of
waters O
>-xJ
sulphate >40 <10 Ca+2 less than 4.5 in all waters
Mg+2 less than 4.0 in all waters O
F
>-xJ
•....•
Chloride III ehloride-biearbonate 30-15 >20 Na + + K+ prevail in all types of trl
r-
waters t:J
IV ehloride-sulphate 15- 5 >20 <25 Ca+2 less than 12.5 in the water
::s
of high sa line facies
V ehloride < 5 >40 <10 Mg+2 less than 6.0 in all types ~ >
>-3
waters trl
~
en
CHEBOTAREV'S CLASSIFICATIO 265

geochemical groups of waters. The three major groups of waters are divided
into genetic types, which are determined from the absolute concentrations
of the dissolved constituents expressed in reacting values in percent. The
bicarbonate group contains three genetic types of water, namely: (1) bicar-
bonate; (2) bicarbonate-chloride ; and (3) chloride-bicarbonate. The amounts
of bicarbonate plus carbonate and chloride plus sulfate determines the
genetic type. The sulfate group is subdivided into two genetic types: (1)
sulfate-chloride; and (2) sulfate. The chloride group is divided into three
genetic types: (1) chloride-bicarbonate; (2) chloride-sulfate ; and (3) chloride.
Chebotarev relates the genetic types to the products of weathering because
he believes that although the concentration of dissolved solids in subsurface
waters may vary substantially, the types of soluble salts remain largely
unchangeable.
The water classes are related to the products of weathering, rainfall, and
drainage conditions. Class I corresponds to soluble products from the weath-
ering of orthoeluvium or igneous and highly metamorphosed rocks and their
silicate compounds. Class II waters are related to products of weathering
from the same silicates and calcareous accumulations. Class III waters
primarily are related to weathered products from calcareous accumulations.
Class IV waters are related to weathering of alluvial, detrital, and gypsum
deposits. Class V waters are related to marine deposits plus weathering of the
products that derived the Class IV waters.
Table 8.1V shows the approximate reacting values in percent for waters
found in some oilfields (Chebotarev, 1955). For example, in such waters in
the bicarbonate group, the major cations are sodium plus potassium with the
reacting value percentage for calciurn and magnesium less than 2.5 if the
saline facies are high. (The divisions between his saline facies are shown in
Table 8.V.)
Table 8.V shows the relationships of hydrodynamic zones to the geo-
chemistry of the water and the geological environment. The zones are: (1)
recharge with active water exchange; (2) pressure with delayed water
exchange; and (3) accumulation with stagnant conditions.
The equilibrium of the chemical systems (those typical of the major geo-
chemical group of waters) is a criterion called the coefficient of water
exchange (Ke) and is computed as:
Na(K)HC03 + (Ca,Mg,) (HC03 b
K =--------~.~--~--~~-=~--~~-----
e Na(K)CI + (Ca,Mg)CI2 + Na2804 + (Ca,Mg)804
The absolute and relative coefficients of water exchange for the three major
groups of waters are as follows:

«; (absolute) K; (relative )
Bicarbonate waters 1.55 96.9
8ulfate waters 0.11 6.9
Chloride waters 0.016 1.0
t-:)
TABLE 8.V
en
en
Relationships of h ydrodynamic zones to the geochemistry of water and the geological evironment (after Chebotarev, 1955)

Hydrodynamic zone Geochemistry of water Geological environment

recharge- water- class hydro-chemical approximate cornmon structures relation to water depth exa~ples
discharge exchange facies salinity terms for (ft)
cycle (ppm) water

Zone of active 1 and II low saline facies 180-2,400 fresh different intensive f1ush usually less eyerywhere
recharge exchange (sorne- than 500
times
IJI)

transitional 2,400-11,400 brackish deep portions of delayed f1ush sometimes Great Artesian Basin ,
(typical) facies structures with 5,000-7,000 Rocky Mountain oil-
peculiar geo- field and others
chemical en-
vironment

high saline facies 11 ,400-37 ,800 saline hampered f1ush

Zone of delayed III and low saline facies 400-2,500 fresh differenl inadequate f1ush usualíy less everywhere
o
r-
pressure exchange IV than 1,000 >

--
ir:
trans .•...
ional 2,500-7,400 brackish
tr:
deeper porti ons circulation and sometimes pre-Ca ucasia n Basin,
'Tj
(typical) facies of structures, drainage Iimited 3,000-4,000 South Dakota Basin,
folded zo nes some oilfield areas o
>

Zone of stagna nt V
high saline facies

low saline facies


7,400-19,300

1,500-20,000
saline

fresh and differenl salt accumulation different chiefly in arid regions;


-
>-3
O
Z
accumulation condition saline prevails upon deeper portions o f many O
'Tj

--
leaching artesian basins: some
oilfields O
r-
transitional 20,000-90,000 saline and deeper portions sometimes 'Tj
(typical) facies brines of structures, 8,000-13,000 trl
highly folded r-
zones tj
::e
high saline facies 90,000-300,000 brines water exchange many oilfield areas
manifests on geo- (Louisiana, Alberta, etc.) ~
logical scale time trl
::<l
r:n
SCHOELLER'S SYSTEM 267

The absolute and relative coefficients were derived from assumed chemical
compositions of typical waters of the major geochemical groups. The abso-
lute value can be not lower for the group, but it can be relatively higher.
Several thousand analyses of waters associated with oil pools in the world
were used to formulate the typical waters. As the water-exchange conditions
deteriorate, the type of water changes, and the changes are related to altera-
tion or diagenesis of the waters (metamorphism). The data in Table 8.V give
hydrochemical facies, common names of various waters as related to
dissolved solids concentrations, geological structures, flush or water circula-
tion, depths, and examples where some of the water types are found.
The highly concentrated chloride waters are primarily associated with oil
occurrences; however, this is not always true. A prime determinant of the
chemical composition of oilfield waters is the hydrodynamic situation and
the type of trapo For example; an intensively flushed zone will contain a
different type of water from a zone with limited circulation. The type of
basin strongly influences the type of water that. is likely to be found. For
example, an open basin probably will contain artesian waters of the bicar-
bonate group, a partly closed basin may contain artesian or subartesian
waters of the bicarbonate or sulfate groups, while a closed basin is more
likely to contain bicarbonate waters on the flanks of the basin with sulfate
and chloride water s in the deeper areas.
Chebotarev (1955) applied his classification to 917 subsurface waters in
oilfields in the world. The classification indicated that 73.7% of the waters
were of the chloride genetic type, 23.0% of the bicarbonate type, and 3.3%
of the sulfate type. Most of the sulfate and bicarbonate types were found in
the Rocky Mountain areas of the United States and probably were mixtures
containing infiltrating meteoric water.

Schoeller's system

A French geochemist, Schoeller (1955), classified water s on the basis of


their dissolved constituents and in the following order of importance: (1)
chloride; (2) sulfate; (3) bicarbonate plus carbonate; (4) index of base
exchange (IBE); and (5) relationships of anions to cations. This system
separates waters into six primary types based upon their amounts of dis-
solved chloride and four subgroups based on their concentrations of sulfate.
The amounts of bicarbonate and carbonate ions give additional differentia-
tion, and an index of base exchange indicates exchange of ions in the waters
with ions in associated clays. Table 8.VI outlines Schoeller's classification.
As shown in Table 8.VI the chloride concentration separates waters into
six types. Waters from several oil-producing regions were classified and the
sequence Cl > S04 > HC03 was found to occur in very high chloride waters
and in sea waters, especially when they are saturated with CaS04' If the
water s are not saturated in CaS04' the sequence Cl" > HC03 - > S04 -2 .is
predominant, and in low-chloride waters the predorninant sequence is
HC03 - > Cl" > S04 -2 .
268 CLASSIFICATION OF OILFIELD WATERS

TABLE 8.VI

Schoeller's scheme for classifying petroleum reservo ir waters*

Chloride concentration as Cl

Very high if >


700
Marine ir 420 - 700
High if 140 - 420
Average if 40 - 140
Lowif10-40
Normal if < 10

Sulfate concentration as S04 -2

Very high if > 58


High if 24 - 58
Average if 6 - 24
Normal if< 6
N ear sa t ura tiIon w hen : '(80
en V 4 2) (Ca+2) > 70

2
Bicarbonate plus carbonate concentrotion as HC03 - + C03-

High if 7>
Normal if 2 - 7
Low if < 2
However, he recommends using ~(HC03 + C03 2)2 (Ca+2) rather
than HC03 - + C03-2

lndex of base exchange (IBE)

If Cl" > Na+ then IBE = (Cf" - Na+)/Cl-


If Na+ > CI- then IBE = (CI- - Na+)/(804 -2 + HC03 - + C03 -2)

lmportance o] anions and cations

CI- > 804 > C03-


-2
2

CI- > C03- > 80


2
4-
2

C03 -2 > cr > 804-2


C03 -2 > 804-2 > cr
Na+ > Mg+ > Ca+
2 2

Na+ > Ca+ 2> Mg+2


* All constituents are calculated in epm.
SCHOELLER'S SYSTEM 269

Also, in very high chloride waters only the sequence Na" > Ca+2 > Mg+2
is found. As the Cl" decreases, the sequence Na+ > Mg+2 > Ca+2 becomes
more frequent. In very high chloride waters S04 -2 > Ca+2, but in less
concentrated waters the opposite may occur. The sequence HC03 - < Ca+2
always is found in very high chloride waters. In less concentrated chloride
waters, either HC03 - < Ca+2 or HC03 - > Ca+2 may be found, while in
low chloride water s HC03 - > Ca+2 is predominant.
Schoeller used an arbitrary value of 70 for ~r:S::-C0:::-4---="2
-x---:C==-a-+"'2
to indicate
that a water is saturated with CaS04. (This is not necessarily true because
some waters, depending upon their other dissolved constituents, can contain
smaller or larger amounts.) He divided waters into four additional types
depending upon their amounts of sulfate,
Saturation with CaS04 was found to occur only in very high chloride
waters. The calcium concentration always is very high - ranging from 150 to
1,100 epm - in high chloride water s which have S04 -2 > 58 epm and
usually is less than 150 epm in high chloride waters where S04 -2 < 58 epm.
All petroleum waters even if saturated in CaS04 have a low S04/Cl ratio
which is attributed to reduction of sulfates and high concentrations of
chloride. The ratio never exceeds one except in low or normal chloride
waters.
The third subgroup contains three additional types depending upon the
amounts of bicarbonate and carbonate in the waters. The preferred formula
for this calculation is 1 (HC03 + C03 2)2 (Ca" 2) which is proportional to
the gaseous pressure of CO2 in equilibrium with CaC03 in the water. As the
Cl" increases, the tendency is for Ca" " to increase and HC03 - to decrease;
however, because the Ca+2 increases, the product of 1(HC03 + C03 2)2
(Ca+2) does not vary greatly.
As the waters move in their subsurface environment their dissolved ions
. have a tendency to exchange with those in the rocks. Two extreme types of
adsorption can be noted in addition to intermediate types of adsorption. The
extreme types are a physical adsorption or the Van der Waals adsorption with
weak bonding between the adsorbent and the constituent adsorbed and a
chemical adsorption with strong valence bonds. Both of these adsorptions
can act simultaneously.
Cations can be fixed at the surface and in the interior of the associated
minerals. These fixed cations can exchange with the cations in the water.
When the exchange occurs, there is an exchange of bases. With the right
physical conditions of the adsorbent, similar exchange can occur with the
anions. Some of the formation constituents that are capable of exchange and
adsorption are argillaceous minerals, zeolites, ferric hydroxide, and certain
organic compounds.
Particle size influences rates and capacities, if the solids are clays such as
illite and kaolinite. The.rate increases with decreasing particle size. However,
if a larger mineral has a lattice, the exchange can easily occur on the plates.
270 ' CLASSIFICATION OF OILFIELD WATERS

The concentration of exchangeable ions in the adsorbent and in the water is


important. More exchange will occur when the solution is highly concen-
trated.
8choeller (1955) usedthe formula:

(a - x) =- (_x_)
K
a-x
't»

to indicate the relationship that exists between the initial concentration, a,


of the cations in milliequivalents in the unreacted water, and x, which equals
the final concentration of the cations in milliequivalents in the water after
equilibrium or reaction with the rocks. The amounts of cations exchanged
by passing from the liquid to the rock or clay is a - x and the index of base
exchange (lEE) =- (a x s]«. By substitution:
+

lEE =- !i(_x_)
a a-x
I/p

The IBE is used to indicate the ratio between the exchanged ion s and the
same ions as they originally existed. For example, assume that in the original
water there were as many equivalents of Cl" as (Na" + K+), and that when
the Na" and K+ of the water exchanged with the alkaline earths in the rocks
alkaline exchange occurred, then:
_ Cl- - (Na+ + K+)
lEE - Cl

and this value is positive if the equivalents are Cl" > (Na" + K+). Theoret-
ically all the halides should be included as Cl" and all the alkalies as Na" or
(Na" + K+).
However, when the alkaline earth ions in the water exchange for alkali
metal ions on the rocks then:
lEE =- Cl- - (Na+ + K+)
804 -2 + HC03 - + N03 -

and this value is negative if the equivalents are Cl" < (Na" + K+). The lack
of equilibrium between the halides and the alkalies is not always a charac-
teristic of base exchange because sea water has a positive value without the
occurrence of base exchange. Negative values usually are observed for water
coming from altered crystalline rocks. Waters with an IBE equal to or greater
than 0.129 can be true connate petroleum reservo ir waters. Waters with a
negative IBE are waters of meteoric origin that have infiltrated into marine
sediments.
Comparison of petroleum-reservoir waters with other types of subsurface
water s revealed that the other waters have most of the same characteristics
but generally have a much higher 804 -2 concentration and a lower
..y(HC03 )2 (Ca+2) or Kr. Waters that are in contact with organic matter
SCHOELLER'S SYSTEM 271

(other than petroleum), such as bitumens, lignites, and coals, resemble petro-
leum reservoir water, but the frequency of a Kr above normal is greater in
petroleurn-associated waters. Waters related to magmatic reactions com-
monly possess high concentrations of HC03-.
Schoeller's (1955) study of petroleum reservo ir waters indicated that a
positive IBE is more frequent as the Cl" increases. A negative lEE is more
frequent as the Cl" decreases, and a negative value is predominant in low
and normal chloride waters associated with petroleum. In fact, this charac-
teristic appears specific for petroleum reservoir waters since in other subsur-
face waters a positive index occurs as frequently as a negative index.
Ancient sea water (connate water) deposited with the sediments usually
has an lEE > 0.129 and a Cl/Na > 1.17. lntruding meteoric water in sedi-
mentary marine rack has an IBE < 0.129 and Cl/Na < 1.17. Petroleum-
reservoir waters with an IBE greater than sea water 0.129 also have the
characteristics Cl/Na > 1.17, Cl/ea < 26.8, Cl/Mg > 5.13, Mg/Ca < 5.24; a
very high value for.(/ (HC03 )'{(Ca+2) indicating sulfate reduction; low con-
centrations of HC03 -; and frequent high concentrations of NH4 +.
Petroleum-reservoir waters containing infiltrating meteoric water mixed
with ancient sea water have an IBE less than sea water, 0.129, and the
characteristics Cl/Na < 1.17, the ratio Mg/Ca increases and appraaches but
never equals 5.24, and the ratios Na/Ca and Na/Mg decrease as the dissolved
salid s increase.

Gases in petroleum-reseruoir waters

Schoeller (1955) noted that there should be equilibrium between the free
petraleum gases and those dissolved in the water. Considering the solubility
of the gases, those in the water should reflect the composition of the petra-
leum. Components characteristic of petraleum accumulations are ethane,
prapane, butane, pentane, ethylene, and prapylene. Associated components,
which may also be present in volcanic waters and in waters in contact with
other organic matter such as coal, peat, and lignite as well as in petroleum-
associated waters, are favorable components. These are methane, carbon
dioxide, organic nitragen, hydragen sulfide, helium, radon, and the absence
of oxygen. Other components or universal components found in all types of
waters are nitrogen and argon.
The top water s can contain gases such as H2 S, CO2 , and CH4 but beca use
theydo not contact the petroleum deposit they are not similar in com-
position to the bottom waters. The edge waters are in contact with the
petroleum and are characterized by higher amounts of HC03, sulfate reduc-
tion, and the presence of H2 S, NH4, and small amounts of dissolved hydra-
carbons.
272 CLASSIFICATION OF OILFIELD WATERS

Oilfield brine analyses

About 4,000 oilfield brine analyses were classified. AJI of these analyses
are now in the U.S. Bureau of Mines (1965) open-file report on oilfield
brines. The data are on automatic data processing magnetic tape as well as in
afile of computer printout sheets listed by State, county , sedimentary basin,
formation, etc.
These brine data were collected by the Bureau of Mines because the value
of oilfield brine analyses in the study of various petroleum-related problems
was recognized early in the history of petroleum and natural gas. Before
1928 the Bureau of Mines had indicated in several reports (Ambrose, 1921;
Swigart and Schwarzenbek, 1921; Collom, 1922; MilIs, 1925; Reistle, 1927)
ways in which the analyses could be used. In earlier years, confusion existed
because of greatly varying methods of analysis that were used. A paper
presenting the methods used by the Bureau of Mines was published by
Reistle and Lane (1928). This system of determining the characteristic con-
stituents of oilfield waters and of calculating and reporting results was
widely adopted 9Y the petroleum industry. Later the Bureau of Mines
cooperated with several interested agencies, and a more detailed report with
more modern methods of analyzing oilfield water s was published (American
Petroleum Institute, 1968).
The Bureau of Mines has an oilfield water analysis laboratory at the
Bartlesville Energy Research Center, Bartlesville, Oklahoma. These 4,000
samples were analyzed at this laboratory.

Analysis methods

The specific gravity of each sample is determined so that a correct aliquot


size can be taken for a specific ion analysis. Chloride is determined by
titration with silver nitrate, carbonate and bicarbonate are determined by
titration with a standard acid, and a pH meter is used to determine the end
points. This alkalinity determination should be completed at the time of
sampling for accurate results. However, most of the data for the 4,000
samples were obtained by analysis in the laboratory and were completed
within 6 to more than 48 hours after sampling. Therefore, the alkalinity data
cannot be considered absolute but only relative.
The calcium was determined by titration of calcium oxalate with perman-
ganate until about 1957, about which time it was determined by com-
plexometric titration such as with disodium ethylenediametetraacetic acid
(EDTA) until about 1969; since then, it has been determined by atomic
absorption. Magnesium determination has a similar history. It was precipi-
tated as the pyrophosphate until about 1957, and titrated with EDTA until
1969, from then to now it has been determined by atomic absorption.
Sulfate was determined by precipitation as barium sulfate, and this meth-
od stiJl is used. Sodium was determined by calculation from the difference
OILFIELD BRINE ANALYSES 273

between the reacting values of the assumed total anions and cations until
about 1960, after which time it was determined by flame photometry or
atomic absorption.
Several other analyses for dissolved constituents in oilfield brines are now
made by the Bureau of Mines. For example, potassium, lithium, rubidium,
and cesium are determined by atomic absorption or flame photometry;
strontium and barium by atomic absorption; manganese, iron, and boron by
atomic absorption or titrimetric methods; and bromide and iodide by
titrimetric methods.
The precision of the methods is as follows: alkalinity, 2-3% of the
amount present; sodium, 2-5% of the amount present; calcium and magne-
sium, 4-5% of the amount present; sulfate, 1-2% of the amount present;
chloride, 1 % of the amount presento If sodium is calculated, the precision
value reflects the sum of the precision data for the data from which it is
calculated plus the undetermined dissolved constituents. The significant
figures for the analytical data are all the certain digits and only the first
doubtful digit. This number usually is limited to three significant figures
except for specific gravity, where four or five are common.
It often is recognized that the sampling metbod is as important as the
analytical method. This certainly is true of oilfield brines.

Field sampling methods

Most of the 4,000 samples were obtained only from wells wbere reason-
able assurance was evident tbat the formation brine was not contaminated
by drilling fluids or by intrusion of water from other formations. Wells were
selected on the basis of age, type of completion, and production of fluids.
Samples were not taken from some gasfields because of tbe likelihood of
dilution by water condensed from vapor carried up the hole witb tbe gas.
Some samples were taken from gas-condensate wells tbat produced large
volumes of brine.
In many cases the electrical resistivity measurement was made on tbe
sample at tbe time tbe sample was taken, and resulting values were compared
with measurements from other samples from the same formation witbin tbat
field or nearby fields. Obvious discrepancies were eliminated by sampling
additional wells.
Nearly all samples were withdrawn at production wellbeads, and tbe water
was separated from the oil in portable separators. A few samples were taken
of brines from formations that did not produce enough water to permit
taking samples at individual wellbeads. Such samples were taken from gun-
barreIs or oil-water separators.
Samples were taken in clean, l-gallon glass jugs that were first rinsed
several times with the water sampled and then filled, capped, and labeled. In
a few instances samples were obtained by the producer from comparatively
isolated small pools or fields and shipped to the laboratory in I-gallon poly-
etbylene jugs.
274 CLASSIFICATION OF OILFIELD WATERS

Application of the classification systems

Several investigators have applied the classification systems to determine


their usefulness in studies related to exploration and production of petro-
leum. Ostroff (1967) concluded that Palmer's system is less useful than the
systems of Sulin and Schoeller. Further he concluded that the Sulin system
is more applicable to petroleum formation waters because many such waters
contain more than 2,000 epm of dissolved solids, and the Schoeller system
tends to lump these highly concentrated waters together. The index of base
exchange (IBE) in Schoeller's system however a pears to have merit for
certain interpretations as does the Ca+2 x S04 -2.
Dickey (1966) concluded that the Palmer system does not correlate very
well with the geology of oil reservoirs, that Schoeller's nomograph is useful,
but that the Sulin system appears to conform better with geology. He also
noted that relating water types to geology and flow patterns should be useful
in exploration and that distinguishing between a stagnant water and artesian
related waters should be highly significant in a new oilfield development
area.
The Sulin system appears to be more generally applicable than the other
systems in studies of water s from petroleum reservoirs. Because of this and
because the Schoeller system appears to have merit in studying certain types
of oilfield waters, it was decided to apply portions of the Sulin system, the
Schoeller system, arid Bojarski's modification of the Sulin system to a study
of brines from various sedimentary basins of the United States that are
known to be related to petroleum and natural gas.

Calculations

The analyses of most oilfield waters are reported in units of milligrams per
liter (mg/I). The conversion of mg/l to equivalents per million (epm) is done
using the following formula:
mg/l of ion _.
. . ht f i - epm l Ofl
T it fb . at omlC welg o lOn
speci ICgravi y o r..¡nex valence of ion

Table 8.VII provides formulas for calculating the epm for many of the
common constituents found in oilfield brines. If the constituent is reported
in parts per rnillion (ppm), it is not necessary to divide by the specific gravity
of the brine.
The sum of the epm's (L epm) shown in Table 8.VII are converted to
Lr/100 g of water for the Sulin calculations by moving the decimal to the
left in the L epm and calling this Lr. The term s is used to indicate the
percent of equivalent of a given constituent. The percentage equivalent (s) of
each ion is determined by dividing the equivalents per 100 g of water by the
total equivalent in 100 g of water. For example, if the r for sodium equals
APPLICATION OF THE CLA88IFICATION 8Y8TEM8 275

TABLE 8.VII

Formulas for converting milligrams per liter to equivalents per million for constituents
commonly found in oilfield waters

Lithium mg/l Li+ x O.1442/sp. gr.* = epm i.r'


Potassium mg/l K+ x O.0256/sp. gr. = epm K+
Sodium mg/l Na+ x O.0435/sp. gr. = epm Na+
Magnesium mg/l Mg+2 x O.0823/sp. gr. = epm Mg+2
Calcium mg/l Ca+2 x O.0499/sp. gr. = epm Ca+2
8trontium mg/l 8r+2 x O.0228/sp. gr. = epm 8r+2
Barium mg/l Ba +2 x O.0146/sp. gr. = epm Ba+2
Carbonate mg/l C03-2 x O.0333/sp. gr. = epm C03 -2
Bicarbonate mg/l HC03 x O.0164/sp. gr. = epm HC03-
8ulfate mg/1804-2 x O.0208/sp. gr. = epm 804 -2
Chloride mg/l Cl- x O.0282/sp. gr. = epm Cl"
Bromide mg/l Br- x O.0125/sp. gr. = epm Br"
Iodide mg/t I" x O.0079/sp. gr. = epm 1

~epm

* 8pecific gravity.

200.6 and the ~r for the total equivalents equals 518.8, then
200.6/518.8 x 100 = 38.7, or 38.7 percentage equivalents for sodium.
The Sulin classification considers only the macro constituents; if ion s such
as potassium, lithium, strontium, barium, bromide, and iodide are deter-
mined, they should be added to their associated macro constituents to be
properly considered in the analysis reporto For example, when sodium,
potassium, and lithium are determined by atomic absorption the total mg/I
of each is reported. Therefore, the epm Na + epm K +epm Li should be
added to obtain the correct r value for sodium:
N epm Na epm K epm Li
r a= 10 + -W + 10

epm ea epm Sr epm Ba


r e a= +---+
10 10 10

- epm el epm Br epm I


r
el - 10 + 10 + 10
The r values then are divided by the ~r and multiplied by 100 to obtain
the s value or percentage equivalents. The s values are used to determine the
type, class, and other Sulin values of the water as illustrated in Table 8.1,
where a = sNa, b = sea + sMg, and d = sel + SS04. The epm values are used
to determine the Schoeller characteristics such as the degree of chloridiza-
tion, the degree of sulfation, lEE, etc., illustrated in Table 8.VI.
276 CLASSIFICATION OF OILFIELD WATERS

TABLE 8.VIII

Classification of some oilfield waters from 10 formations in eight sedimentary basins

No. State Formation Basin Depth Concentration (epm)


(m)
Na+ Mg+2 Ca+2 HC03 - S04-2 CI-

1 Kans. Arbuckle Cent.Kans. 1,050 634.6 60.6 133.0 7.7 49.0 773.8
2 Kans. Arbuckle Cent.Kans. 1,091 581.6 58.4 121.8 6.5 22.7 732.3
3 Kans. Arbuckle Cent.Kans. 1,023 282.0 24.8 40.2 9.5 22.1 313.9
4 Kans. Arbuckle Cent.Kans. 1,102 639.9 56.6 13.0 2.2 46.6 827.0
5 Kans. Arbuckle Cent.Kans. 992 430.3 38.8 65.4 8.5 17.5 510.9
6 Kans. Arbuckle Cent.Kans. 1,152 458.7 34.4 81.3 3.4 45.9 523.2
7 Kans. Arbuckle Cent.Kans. 1,174 473.2 53.1 119.4 5.1 36.0 705.0
8 Kans. Arbuckle Cent.Kans. 949 356.4 57.4 100.4 12.8 39.9 456.7
9 Kans. Arbuckle Cent.Kans. 1,195 446.4 35.4 71.2 9.0 30.0 512.7
10 Kans. Arbuckle Cent.Kans. 1,104 577.4 49.4 112.1 35.9 45.5 685.8
11 Kans. Lansing Cent.Kans. 928 1,759.6 266.6 373.9 1.8 0.0 2,398.9
12 Kans. Lansing Cent.Kans. 1,075 570.4 117.4 174.6 2.4 34.9 826.6
13 Kans. Lansing Cent.Kans. 966 1,899.9 266.4 442.1 0.3 0.8 2,605.G I
14 Kans. Lansing Cent.Kans. 999 1,728.2 232.0 512.4 1.7 2.5 2,469.4
15 Kans. Lansing Cent.Kans. 902 1,903.8 274.5 478.8 1.1 0.0 2,655.6
16 Kans. Lansing Cent.Kans. 1,009 2,514.8 309.3 532.7 0.4 1.5 3,353.6
17 Kans. Lansing Cent.Kans. 1,063 1,854.7 225.2 449.8 0.5 1.1 2,529.5
18 Kans. Lansing Cent.Kans. 1,l48 2,087.2 223.6 384.8 1.4 1.0 2,706.1
19 Kans. Lansing Cent.Kans. 1,172 2,414.5 251.0 519.0 0.7 2.4 3,179.9
20 Kans. Lansing Cent.Kans. 853 1,898.4 210.8 416.7 0.8 0.7 2,515.5
21 Okla. Wilcox Cherokee 1,228 2,366.5 193.3 810.1 1.1 7.7 3,360.6
22 Okla. Wilcox Cherokee 1,731 2,188.4 209.5 529.4 0.3 9.8 2,917.2
23 Okla. WilcOx Cherokee 904 2,161.5 163.2 530.3 0.4 7.3 2,847.4
24 Okla. Wilcox Cherokee 1,539 2,365.3 200.8 445.6 1.1 15.9 2,992.0
25 Okla. Wilcox Cherokee 1,106 1,997.1 140.4 457;3 0.9 7.8 2,584.3
26 Okla. Wilcox Cherokee 1,020 1,990.0 161.3 513.7 0.7 13 ..3 2,650.8
27 Okla. Wilcox Cherokee 582 1,587.2 151.2 287.5 1.8 1.4 2,021.6
28 Okla. Wilcox Cherokee 1,432 2,459.0 174.5 511.6 1.1 6.0 3,138.8
29 Okla. Wilcox Cherokee 1,972 2,701.4 153.1 606.0 0.7 8.4 3,449.1
30 Okla. Wilcox Cherokee 1,865 2,635.5 180.6 629.6 1.1 8.6 3,436.6
31 Ark. Nacatoch E.Texas 360 233.4 9.0 22.4 3.7 0.0 261.1
32 Ark. Nacatoch E.Texas 465 463.4 39.5 59.5 1.4 0.0 559.5
33 Ark. Nacatoch E.Texas 373 300.4 25.9 36.8 4.3 1.0 357.7
34 Texas Nacatoch E.Texas 905 788.2 9.4 28.7 3.1 0.0 824.6
35 Texas Nacatoch E.Texas 701 481.8 7.8 17.3 3.3 0.0 503.5
36 Texas Nacatoch E.Texas 242 274.0 8.1 14.7 2.5 0.8 294.2
37 Texas Nacatoch E.Texas 650 492.7 10.4 16.5 18.9 0.1 498.4
38 Texas Nacatoch E.Texas 283 295.4 8.1 14.7 2.4 0.9 313.6
39 Texas Nacatoch E.Texas 191 451.5 17.6 44.7 7.1 0.4 506.2.
40 Texas Nacatoch E.Texas 181 490.6 3.6 7.8 2.8 0.6 498.' t

>
* Chloride (epm): (VH)C = 700, (MC) = 420---700, (H)C = 140-420, (A)C = 40-140,
(L)C = 10---40, (N)C = < 10. Sulfate (epm): (VH) = >
58, (H) = 24-58, (A) = 6-24, (N)
=< 6.
APPLICATION OF THE CLASSIFICATION SYSTEMS 277

Sulin Schoeller* y'Ca+2 + S04 2


{/(HC03 + CO2 2)2(Ca+2)IBE

~epm type class Cl- S04 -2

3.8 1,659.0 Cl-Ca SrS2A2 (VH)C (H) 80.7 20.0 0.18


2.3 1,523.6 Cl-Ca SIS2A2 (VH)C (A) 52.6 17.4 0.21
309 692.6 Cl-Ca SI S2A2 (H)C (A) 29.8 15.4 0.10
7.0 1,585.4 Cl-Ca SIS2S3 (VH)C (H) 77.8 8.7 0.23
0.9 1,071.6 Cl-Ca SI S2A2 (M)C (A) 33.8 16.8 0.16
3.2 1,14 7.2 Cl-Ca SI S2A2 (M)C (H) 61.1 9.9 0.12
5.0 1,492.0 Cl-Ca SI S2 A2 (VH)C (H) 65.6 14.6 0.19
6.7 1,024.0 Cl-Ca SI S2A2 (M)C (H) 62.3 25.2 0.22
2.7 1,105.0 Cl-Ca S¡ S2A2 (M)C (H) 46.2 17.9 0.13
5.8 1,506.4 Cl-Ca SI S2 A2 (M)C (H) 71.4 52.5 0.16
8.9 4,800.9 Cl-Ca SI S2 A2 (VH)C (N) 0.0 10.7 0.27
6.6 1,726.6 Cl-Ca SIS2A2 (VH)C (H) 78.0 10.2 0.31
5.C I 5,214.7 Cl-Ca SI S2 A2 (VH)C (N) 19.4 3.5 0.27
9.4 4,946.5 Cl-Ca SI S2 A2 (VH)C (N) 36.0 11.6 0.30
5.6 5,314.0 Cl-Ca SI S2 A2 (VH)C (N) 0.0 8.5 0.28
3.6 6,712.5 Cl-Ca SI S2A2 (VH)C (N) 28.7 4.4 0.25
9.5 5,061.1 Cl-Ca SI S2 S3 (VH)C (N) 22.5 5.1 0.27
6.1 5,404.4 Cl-Ca S¡S2S3 (VH)C (N) 20.4 9.1 0.23
9.9 6,367.7 Cl-Ca SIS2A2 (VH)C (N) 35.9 6.3 0.24
5.5 5,043.1 Cl-Ca SIS2A2 (VH)C (N) 17.8 6.6 0.25
,0.6 6,739.4 Cl-Ca S¡ S2 A2 (VH)C (A) 79.1 9.9 0.30
7.2 5,854.9 Cl-Ca SI S2A2 (VH)C (A) 72.3 3.6 0.25
7.4 5,710.4 Cl-Ca S¡S2A2 (VH)C (A) 62.3 4.7 0.24
2.0 6,021.0 Cl-Ca S¡S2A2 (VH)C (A) 84.3 8.5 0.21
,4.3 i 5,188.0 Cl-Ca S¡S2A2 (VH)C (A) 59.9 7.1 0.23
0.8 5,330.0 Cl-Ca S¡ S2 A2 (VH)C (A) 82.7 6.4 0.25
:1.6 4,051.1 Cl-Ca SI S2A2 (VH)C (N) 20.5 9.9 0.21
,8.8 6,291.3 Cl-Ca S¡ S2A2 (VH)C (A) 55.7 8.7 0.22
.9.1 6,918.8 Cl-Ca S¡ S2A2 (VH)C (A) 71.4 7.1 0.22
;6.6 6,892.3 Cl-Ca S¡S2A2 (VH)C (A) 73.6 9.4 0.23
;1.1 529.8 Cl-Ca SIS2A2 (H)C (N) 0.6 6.8 0.11
,9.5 1,123.5 Cl-Ca S¡ S2 A2 (M)C (N) 0.0 4.9 0.17
,7.7 726.2 Cl-Ca SI S2A2 (H)C (N) 6.0 8.8 0.16
:4.6 1,654.2 Cl-Ca SIS2A2 (VH)C (N) 0.0 6.5 0.04
13.5 1.013.8 Cl-Ca SI S2A2 (M)C (N) 0.0 5.7 0.04
14.2 594.5 Cl-Ca SI S2A2 (H)C (N) 3.6 4.6 0.07
18.4 1,037.4 Cl-Mg S¡A2S2 (M)C (N) 1.6 18.1 0.01
.3.6 635.2 Cl-Ca SI S2A2 (H)C (N) 3.6 4.4 0.06
16.2 , 1.027.8 Cl-Ca S¡ S2A2 (M)C (N) 4.5 13.2 0.11
18.' 1,004.0 Cl-Ca SIS2A2 (M)C (N) 2.2 4.0 0.02
278 CLASSIFICATION OF OILFIELD WATERS

TABLE 8.VIII (continued)

No. State Formation Basin Depth Concentration (epm)


(m)
Na+ Mg+2 Ca+2 HC03
- S04 -2 Cl-

41 Ark. Paluxy E.Texas 1,115 1,246.8 90.6 254.1 0.4 0.0 1,594.4
42 Ark. Paluxy E.Texas 737 586.0 32.2 86.7 2.8 3.3 699.0
43 Ark. Paluxy E.Texas 1,297 1,310.3 107.1 278.5 2.7 .7 1,692.0
44 Ark. Paluxy E.Texas 884 934.9 64.8 197.6 1.5 1.1 1,194.9
45 Ark. Paluxy E.Texas 1,417 1,507.2 138.6 504.4 0.5 3.3 2,208.7
46 Texas Paluxy E.Texas 2,340 1,642.1 22.5 378.5 0.0 9.3 2,033.8
47 Texas Paluxy E.Texas 2,174 1,515.1 51.1 205.9 4.7 3.7 1,763.1
48 Texas Paluxy E. Texas 1,943 1,495.4 89.0 448.9 2.7 8.2 2,022.2
49 Texas Paluxy E.Texas 1,512 205.3 6.4 10.0 14.9 6.0 202.4
50 Texas Paluxy E.Texas 1,943 1,522.8 90.2 448.1 1.8 8.3 2,050.3
51 Ark. Rodessa E.Texas 1,897 1,971.8 157.2 669.7 0.0 8.0 2,789.9
52 Ark. Rodessa E.Texas 1,241 1,943.8 185.1 563.6 0.9 11.7 2,679.0
53 Ark. Rodessa E.Texas 1,033 1,060.3 108.2 289.8 0.0 21.4 1,436.5
54 Ark. Rodessa E.Texas 711 538.6 35.9 75.0 3.0 2.9 643.4
55 Texas Rodessa E.Texas 2,844 1,772.1 127.3 878.8 1.5 2.8 2,773.2
56 Texas Rodessa E.Texas 2,519 1,861.9 217.4 1,084.7 1.1 2.7 3,165.1
57 Texas Rodessa E.Texas 2,722 2,068.5 148.5 900.9 0.6 4.8 3,111.6
58 Texas Rodessa E.Texas 2,115 1,950.3 139.9 726.1 1.8 8.5 2,802.4
59 Texas Rodessa E.Texas 2,289 2,020.0 141.2 741.8 0.8 4.2 2,909.8
60 Texas Rodessa E.Texas 3,062 1,877.5 92.7 610.0 1.0 5.4 2,576.9
61 Texas Woodbine E.Texas 1,047 1,507.0 44.3 161.4 4.2 3.5 1,705.0
62 Texas Woodbine E.Texas 1,750 1,263.0 33.4 56.3 8.2 1.8 1,342.2
63 Texas Woodbine E.Texas 898 341.8 4.4 9.3 11.3 0.6 343.5
64 Texas Woodbine E.Texas 841 1,060.5 32.4 58.9 0.0 0.2 1,161.4
65 Texas Woodbine E.Texas 1,809 1,271.9 59.6 154.9 1.0 4.9 1,478.2
66 Texas Woodbine E.Texas 1,144 809.1 12.7 61.1 7.3 3.7 873.3
67 Texas Woodbine E.Texas 925 478.8 12.4 23.6 7.0 0.1 507.7
68 Texas Woodbine E.Texas 1,442 1,451.3 26.7 213.3 3.9 0.0 1,685.1
69 Texas Woodbine E.Texas 1,596 1,447.3 38.3 144.1 1.6 0.1 1,629.4
70 Texas Woodbrine E.Texas 1,332 1,268.0 30.2 81.9 8.3 4.0 1,367.4
71 Ala. Eutaw Interior Sal. 1,061 1,124.4 45.0 164.7 2.7 0.0 1,330.1
72 Ala. Eutaw Interior Sal. 972 1,102.9 25.6 160.5 2.0 0.0 1,288.9
73 Ala. Eutaw Interior Sal. 1,060 1,186.7 59.6 164.3 2.9 0.0 1,411.8
74 Miss. Eutaw Interior Sal. 1,444 1,733.6 82.7 287.3 0.9 0.0 2,108.5
75 Miss. Eutaw Interior Sal. 2,263 2,009.8 75.9 306.8 5.2 33.6 2,35'8.5
76 Miss. Eutaw Interior Sal. 1,312 1,572.7 88.9 224.9 0.9 0.0 1,890.2
77 Miss. Eutaw Interior Sal. 2,443 1,721.3 256.7 458.0 0.0 0.0 2,433.4
78 Miss. Eutaw Interior Sal. 1,690 1,925.9 61.6 359.9 3.9 0.0 2,352.4
79 Miss. Eutaw Interior Sal. 1,315 1,579.6 60.6 252.9 2.3 0.0 1,893.5
80 Miss. Eutaw Interior Sal. 2,469 2,153.7 87.7 518.9 0.0 1.0 2,765.6

* Chloride (epm): (VH)C ="> 700, (MC) = 420-700, (H)C = 140--420, (A)C = 40-140,
(L)C = 10--40, (N)C = <
10. Sulfate (epm): (VH) = > 58, (H) = 24-58, (A) = 6-24, (N)
~< 6.
APPLICATION OF THE CLASSIFICATION SYSTEMS 279

Sulin Schoeller* VCa +2 + S04 2 V'(HC03 + CO2 2)2(Ca+2) !BE

2:epm type class CI- S04 -2

3,186.6 CI-Ca SIS2S3 (VH)C (N) 3.1 3.8 0.22


1,410.3 CI-Ca SI S2 A2 (M)C (N) 17.1 8.8 0.16
3,391.5 CI-Ca SI S2 A2 (VH)C (N) 14.1 12.8 0.23
2,395.0 CI-Ca SI S2A2 (VH)C (N) 15.0 7.6 0.22
4,363.0 CI-Ca SI S2A2 (VH)C (N) 41.0 5.4 0.29
4,086.4 CI-Ca SI S2 A2 (VH)C (A) 59.5 0.0 0.19
3,543.9 CI-Ca SI S2 A2 (VH)C (N) 27.9 16.6 0.14
4,066.7 CI-Ca SIS2A2 (VH)C (A) 60.9 13.6 0.26
445.2 S04 -Na SI A2 S2 (M)C (A) 7.7 13.1 0.0
4,121.6 CI-Ca SI S2A2 (VH)C (A) 61.0 11.1 0.26
5,596.8 CI-Ca SI S2 A2 (VH)C (A) 73.3 0.5 0.29
5,384.3 CI-Ca SI S2 A2 (VH)C (A) 81.3 8.0 0.27
2,916.3 CI-Ca SIS2A2 (VH)C (A) 78.7 0.4 0.26
1,299.0 CI-Ca SI S2 A2 (M)C (N) 14.8 8.8 0.16
5,556.0 CI-Ca SI S2 A2 (VH)C (N) 50.1 12.9 0.36
6,333.1 CI-Ca SIS2S3 (VH)C (N) 54.4 11.1 0.41
6,235.2 CI-Ca SIS2A2 (VH)C (N) 66.2 7.3 0.34
5,622.2 CI-Ca SI S2 S3 (VH)C (A) 79.0 13.5 0.30
5,818.1 CI-Ca SIS2S3 (VH)C (N) 55.9 7.7 0.31
5,163.7 CI-Ca SIS2S3 (VH)C (N) 57.4 8.4 0.27
3,425.6 CI-Ca SI S2 A2 (VH)(C) (N) 23.9 14.2 0.12
2,705.0 CI-Ca SI S2 A2 (VH)C (N) 10.0 15.6 0.06
711.0 CI-Mg SIS2AI (H)C (N) 2.4 10.6 0.00
2,313.7 CI-Ca SIS2S3 (VH)C (N) 4.0 0.0 0.09
2,970.7 CI-Ca SI S2A2 (VH)C (N) 27.6 5.4 0.14
1,767.3 CI-Ca SI S2AI (VH)C (N) 15.1 14.8 0.07
1,029.9 CI-Ca S; S2A2 (M)C (N) 1.5 10.5 0.06
3,379.5 CI-Ca SI S2 A2 (VH)C (N) 0.0 14.8 0.14
3,261.0 CI-Ca SI S2 A2 (VH)C (N) 3.7 7.4 0.11
2,760.0 CI-Ca SI S2A2 (VH)C (N) 18.1 17.8 0.07
2,667.1 CI-Ca SI S2 A2 (VH)C (N) 0.1 10.6 0.15
2,580.1 CI-Ca SI S2 A2 (VH)C (N) 0.1 8.8 0.14
2,825.4 Cl-Ca SIS2S3 (VH)C (N) 0.1 11.1 0.16
4,213.1 CI-Ca SI S2 S3 (VH)C (N) 0.1 6.2 0.18
4,789.9 CI-Ca SI S2 A2 (VH)C (H) 101.6 20.3 0.15
3,777.7 CI-Ca SI S2 S3 (VH)C (N) 0.0 5.7 0.17
4,869.6 CI-Ca SI S2 A2 (VH)C (N) 0.0 0.0 0.29
4,704.8 CI-Ca SIS2S3 (VH)C (N) 0.0 17.8 0.18
3,789.1 CI-Ca SI S2 S3 (VH)C (N) 0.0 11.0 0.17
5,527.1 Cl-Ca SI S2 S3 (VH)C (N) 22.7 0.0 0.22
280 CLASSIFICATION OF OILFIELD WATERS

TABLE 8.VIII (continued)

No. State Formation Basin Depth Concentration


(m)
Na+ Mg+2 Ca+2 HC03 - S04-2 CI-

81 Miss. Wilcox Miocene 1,975 2,031.8 36.8 100.3 4.5 0.0 2,163.3
82 Miss. Wilcox Miocene 1,748 1,847.9 48.4 94.9 3.8 0.0 1,985.9
83 Miss. Wilcox Miocene 1,412 1,362.9 15.0 82.5 4.4 0.0 1,455.8
84 Miss. Wilcox Miocene 1,871 2,157.3 54.1 89.9 6.8 0.0 2,114.0

85 Miss. Wilcox Miocene 2,330 1,972.2 65.8 95.7 3.9 7.2 2,122.3
86 Miss. Wilcox Miocene 1,646 1,992.4 38.0 81.9 3.0 0.0 2,108.5
87 Miss. Wilcox Miocene 2,162 2,043.7 25.5 132.2 7.2 0.0 2,194.5
88 Miss. Wilcox Miocene 2,268 2,198.7 83.8 90.7 3.9 0.3 2,369.8
89 Miss. Wilcox Miocene 1,552 1,280.8 36.7 56.3 9.4 0.0 1,364.3
90 Miss. Wilcox Miocene 1,327 1,318.9 20.9 99.6 3.9 0.0 1,434.8
91 La. Smackover N .Louisiana 3,109 1,589.7 86.7 1,256.8 0.0 8.8 2,940.8
92 Ark. Smackover N.Louisiana 2,103 2,444.9 275.9 1,392.8 2.0 4.6 4,105.9
93 Ark. Smackover N.Louisiana 2,399 2,518.1 280.4 1,622.1 1.8 3.7 4,413.8
94 Ark. Smackover N.Louisiana 2,509 2,790.1 222.2 1,641.6 2.3 1.7 4,648.6
95 Ark. Smackover N.Louisiana 2,240 2,418.0 279.5 1,470.8 0.0 4.1 4,163.0
96 Ark. Smackover N.Louisiana 2,526 2,729.0 218.6 1,598.2 2.4 1.7 4,540.4
97 Ark. Smackover N.Louisiana 2,615 4,277.5 10.0 34.9 2.4 0.0 4,312.9
98 La. Smackover N.Louisiana 2,949 2,225.4 149.7 2,282.1 0.0 1.4 4,654.9
99 La. Smackover N.Louisiana 3,271 1,971. 2 173.5 1,668.9 0.0 1.8 3,810.9
100 La. Smackover N .Louisiana 701 681.8 35.2 63.6 5.1 2.1 778.6
101 La. Wilcox W. Gulf 1,399 1,718.5 44.9 90.1 1.4 0.2 1,851.2
102 La. Wilcox W. Gulf 1,814 2,035.8 46.5 124.7 5.6 0.8 2,199.9
103 La. Wilcox W.Gulf 747 1,206.3 43.6 47.2 4.6 0.6 1,291.2
104 La. Wilcox W. Gulf 1,561 1,874.2 40.9 91.0 0.7 0.8 2,003.8
105 La. Wilcox W. Gulf 1,722 2,138.9 41.9 96.2 1.4 1.4 2,273.4
106 La. Wilcox W. Gulf 666 1,057.5 26.6 51.0 0.0 0.0 1,058.4
107 La. Wilcox W. Gulf 1,124 1,379.4 41.8 60.0 2.5 0.6 1,477.6
108 La. Wilcox W.Gulf 2,441 2,119.4 43.7 100.2 6.1 0.0 2,263.4
109 La. Wilcox W. Gulf 2,158 2,132.3 15.3 115.2 4.0 8.0 2,256.4
110 La. Wilcox W. Gulf 471 840.1 26.8 24.1 7.0 0.0 888.4
111 N.D. Silurian Williston 3,633 3,431.2 100.6 993.7 1.5 8.4 4,305.3

* Chloride (epm): (VH)C = >700, (MC) = 420-700, (H)C = 140-420, (A)C = 4040,
(L)C = 10-40, (N)C = < 10. Sulfate (epm): (VH) = >
58, (H) = 24-58, (A) = 6-24, (N)
=< 6.
APPLICATION OF THE CLASSIFICATION SYSTEMS 281

Sulin Schoeller* y'Ca+2 + S04 -2 {!(HC03 - + CO2 -2 )2(Ca+2) !BE

L:epm type class Cl S04 -2

4,336.9 Cl-Ca SIS2A2 (VH)C (N) 0.0 12.7 0.06


3,981.1 Cl-Ca SI S2 A2 (VH)C (N) Q.O 11.2 0.07
2,920.8 Cl-Ca SI S2 A2 (VH)C (N) 0.0 11.8 0.06
4,422.4 HC03 SI A2 Al (VH)C (N) 0.0 16.1 0.0
-Na
4,267.3 Cl-Ca SI S2 A2 (VH)C (A) 26.2 11.4 0.7
4,223.9 Cl-Ca SI S2 A2 (VH)C (N) 0.0 9.0 0.06
4,403.4 Cl-Ca SI S2 A2 (VH)C (N) 0.1 18.9 0.07
4,747.3 Cl-Ca SIS2A2 (VH)C (N) 5.4 11.1 0.07
2,747.7 Cl-Ca SI S2 A2 (VH)C (N) 0.0 17.0 0.06
2,878.2 Cl-Ca SI S2 A2 (VH)C (N) 0.0 11.5 0.08
5,883.0 Cl-Ca SI S2 S3 (VH)C (A) 103.6 0.0 0.46
8,226.2 Cl-Ca SI S2 A2 (VH)C (N) 80.0 17.7 0.40
8,840.1 Cl-Ca SI S2 A2 (VH)C (N) 78.0 17.9 0.43
9,306.8 Cl-Ca SI S2 A2 (VH)C (N) 53.5 21.1 0.40
8,335.6 Cl-Ca SI S2A2 (VH)C (N) 78.0 0.6 0.42
4,454.9 Cl-Ca SI S2 A2 (VH)C (N) 53.4 21.3 0.40
8,637.9 Cl-Ca SI S2 A2 (VH)C (N) 0.0 5.9 0.01
9,313.8 Cl-Ca SI S2 A2 (VH)C (N) 57.5 0.7 0.52
7,626.5 Cl-Ca SI S2A2 (VH)C (N) 55.5 0.6 0.48
1,566.7 Cl-Ca SI S2A2 (VH)C (N) 11.4 11.6 0.12
3,706.6 Cl-Ca SI S2A2 (VH)C (N) 5.1 5.6 0.07
4,413.6 Cl-Ca SI S2 A2 (VH)C (N) 10.2 15.8 0.07
2,593.7 Cl-Ca SI S2A2 (VH)C (N) 5.6 10.1 0.07
4,011.7 Cl-Ca SI S2 A2 (VH)C (N) 8.8 3.6 0.06
4,553.5 Cl-Ca SI S2 A2 (VH)C (N) 11.7 5.9 0.06
2,193.6 Cl-Mg SIA2S2 (VH)C (N) 0.0 0.0 0.0
2,962.2 Cl-Ca SI S2 A2 (VH)C (N) 6.1 7.3 0.07
4,533.4 Cl-Ca SI S2 S3 (VH)C (N) 0.0 15.2 0.06
4,531.3 Cl-Ca SI S2 S3 (VH)C (A) 29.9 12.1 0.05
1,786.7 Cl-Ca SI S2A2 (VH)C (N) 0.0 9.9 0.05
8,840.7 Cl-Ca SI S2A2 (VH)C (A) 90.6 13.0 0.20
282 CLASSIFICATION OF OILFIELD WATERS

In terpretation of the classification

The majority of the 4,000 oilfield waters that were classified fell in the 8 I
82 A2 class and were of the chloride-calcium type. Table 8. VIII illustrates
the classification data for some of the samples from the Arbuckle, and
Lansing Kansas City formations in the Central Kansas Basin; the Wilcox
formation in the Cherokee Basin; the Nacatoch, Paluxy, Rodessa, and
Woodbine formations in the East Texas Basin; the Eutaw formation in the
Interior Salt Basin; the Smackover formation in the North Louisiana Basin;
and the Wilcox formation in the western Gulf Basin.
Most of the samples of the bicarbonate-sodium and sulfate-sodium types
were found at relatively shallow depths. This could indicate that the waters
contained infiltrating water.
The Cl/Na ratio was determined for each sample as suggested by Bojarski.
This ratio was calculated from the epm values, and the ratios ranged from
0.33 to values greater than 1. Values greater than 0.85 are characteristic of
hydrodynamic waters, according to Bojarski (1970),while the more altered
waters found in static environments have ratios of less than 0.50.
Schoeller's classification uses the chloride concentration to separate the
waters into six types. The majority of the 4,000 oilfield waters analyzed for
this study were very high chloride waters (where the chloride epm is equal to
or greater than 700). The sulfate concentration of these waters according to
his classification was not as consistent. Many of them were normal with
sulfate epm less than 6. However, in several waters the sulfate epm was
higher than 24. Few waters contained sulfate in excess of 58 epm.
The JCa+2 x 804 2 exceeded 70 in so me waters, indicating that such
waters were nearly saturated with calcium sulfate. Schoeller did not consider
that the solubility of calcium sulfate increases in the presence of certain
other dissolved ions; therefore the value of 70 may not always indicate
saturation. However, in primary and secondary recovery operations this
value should be considered.
The \y(HC03 + C03 2)2 (Ca+2) was used by Schoeller to determine if a
water was saturated with calcium carbonate, and such a water should have a
value greater than 7. This is not entirely accurate, but the formula does
indicate if the water contains an excess of calcium, which decreases the
carbonate concentration. Many of the 4,000 waters evaluated had values
greater than 7. A distilled water thus saturated would deposit precipitated
calcium carbonate, but the activities of other ions dissolved in a brine cause
the solubility product to be different in the brine.
The predominant cation sequence for these 4,000 oilfield brines was Na"
> Ca+2 > Mg+2. The 804/Cl ratio ranged from 0.00 to 0.34. The ratio 0.34
was fo und :for a bicarbonate-sodium water sample, which may have con-
tained infiltrating water. N one of the chloride-calcium type waters had a
804/Cl ratio greater than 0.17, with many of them 0.00.
The IBE indicates that exchange of metal ions dissolved in the water have
APPLICATION OF THE CLASSIFICATIO SYSTEMS <!83

exchanged with metal ions on the clays. (Schoeller made the arbitrary as-
sumption that the concentrations of sodium and chloride originally present
in the water were equal.) If the lEE is a positive number the exchange was
alkali metals in the water for alkaline earth metal s on the clays, and if the
lEE is negative the exchange was alkaline earth metals in the water for alkali
metal s on the clays. Very few negative numbers were evident when the IBE
was determined on the 4,000 oilfield water analyses. This indicates that most
of the formation waters associated with hydrocarbons had exchanged dis-
solved alkali metals for alkaline earth metal s on the clays. The few samples
that yielded the negative IBE numbers were sulfate-sodium and bicarbonate-
sodium type water s which, according to Sulin, is indicative of terrestrial
derived waters.

Ratios

The ratios Naj(Ca + Mg) and CajMg were calculated from the analytical
data for the 4,000 oilfield water samples. Fig.8.1 illustrates a plot of the
Naj(Ca + Mg) ratio versus dissolved solids for samples from the East Texas
Basin. The ratio tends to decrease with increasing dissolved solids concen-
tration. The depletion of sodium with respect to calcium + magnesium can
be attributed to diagenesis of the waters. This correlates with Schoeller's
lEE, indicating that the alkali metals in the water exchange with alkaline

70

.60 r- • Tertiary
o Upper Cretaceous
•• Lower Cretaceous
50 e--
o Jurassic

-o
o
~ 40 r-
x Pennsylvanian

01

o ~

..
z + 30 -
o
u

20 -
•.
o ••

10 - •• 0•.•.
o o •
x X X • xQ-
o
o 00<0 x X·x ~ •• x~·
~ •••• _ x x

O
1 I I i 1" I ••'" o •• 10 o", 1<> "'e
O 35 70 105 140 175 210 245 280 315 350
OISSOLVEO SOLlOS, gIl

Fig. 8.1. Plot of the Na/(Ca + Mg) ratio versus the concentration of dissolved solids in
some formation waters taken from sedimentary rocks in the East Texas Basin.
284 CLASSIFICATION OF OILFIELD WATERS


7
••
6 •

I~ ,
o + 5
Z o
u 4

3

• •
2 •
• • • •
I
20 80 140 200 260 320
DISSOLVED SOLlDS, gil

Fig. 8.2. Plot of the Na/(Ca + Mg) ratio versus the concentration of dissolved solids in
some formation waters taken from the Rodessa formation in the East Texas Basin. The
line is a least squares fit (y = a + bx + cx2) for the scattered data.

earth metals on the clays, decreasing the dissolved alkali metals and in-
creasing the dissolved alkaline earth metals.
Fig.8.2 shows the scattered data for the brines taken from the Rodessa
formation in the East Texas Basin. The scattered data were submitted to
least squares analysis to determine with more certainty how the ratios varied
with depth and with dissolved solids concentrations. The least squares
formula used was y = a + bx + cx2 or the approximation to a parabola. If the
best fit is a straight line, the solution will regress to it, and this occurred for
the data shown in Fig. 8.2. Also, if this occurs, the value for e-in the formula
is very small. The curve shown in Fig. 8.2 represents a fit of about 88% of
the sum of the mean, a fit of 100% would be ideal.
Preliminary plots of the scattered data indicated that a better fit could be
obtained with the parabolic least squares approximation than with the least
squares approximation for a straight line. A computer was used to obtain the
fit.
The least squares curve indica te s that the concentration of sodium
decreased with respect to calcium plus magnesium until the dissolved solids
were about 210,000 mg/l and then increased. This could be attributed to
exchange of sodium in the water for alkaline earths on the minerals until the
concentration of dissolved solids reached 210,000 mg/l, at which point the
solubility products of the alkaline earths are such that they are unable to
stay in solution. This would correlate with Schoeller's IBE. However, the
IBE does not consider the effect of other ions in solution upon the solubility
product of an ion.
APPLICATION OF THE CLASSIFICATION SYSTEMS 285

KEY

30 - • Tertiary
o Upper Cretaceous
• Lower Cretaceous
o Jurassic
x Pennsylvanian-

-e
o
20 -

°1 ~
u::::!: ••
0<>

• •
10 - • •
o
o • ~ ,. ACto

.. a. X
¡..x
•• l·
X~XX

• •
I I i I 1 1 1 1
°OL----L----L----L----~--~----~--~-----L----L---~
35 70 105 140 175 210 245 280 315 350
DISSOLVED SOLlOS, gil

Fig. 8.3. Plot of the Ca/Mg ratio versus the concentration of dissolved solids in some
formation waters taken from sedimentary rocks in the East Texas Basin.

Fig. 8.3 illustrates a plot of the Ca/Mg ratio versus dissolved solids. Here
the trends appear to be that the Ca/Mg ratio increases as the dissolved solids
increase. This trend also is related to diagenesis of the waters. The calcium
increases as the magnesium decreases, and this may be related to the forma-
tion of dolomite and chlorite, or to reactions with argillaceous minerals.
Fig. 8.4 is a plot of Ca/Mg versus dissolved solids. The Ca/Mg ratios for
some brines from the Rodessa formation in the East Texas Basin were sub-
mitted to least squares analysis, and the results were used to plot Fig. 8.4.
The scattered data in Fig. 8.4 did not yield as good a fit to the formula
y = a + bx + cx2 as did the data in Fig. 8.2, because the fit is about 35% of
the mean where 100% is ideal. The line indicates that the ratio incre~'sed as
the dissolved solids increased. This is the result of magnesium lost from the
brine by reactions to form minerals, ion exchange, or shale membrane filtra-
tion. It is not a result of solubility product because most magnesium com-
pounds are more soluble than calcium compounds.
Fig. 8.5 is a plot of Na/(Ca + Mg) against depth from which the samples
were taken. The trend does not indicate a definite relationship for these
samples.
Fig. 8.6 is a plot of the least squares analysis data for Na/(Ca + Mg) versus
depth for some brines from the Rodessa formation in the East Texas Basin.
Theplot indicates that the sodium concentration decreases with respect to
286 CLASSIFICATION OF OILFIELD WATERS

calcium plus magnesium until the depth is about 2,350 m and then it in-
creases. The curve in Fig. 8.6 represents a 45% mean fit of the scattered data
where an ideal fit is 100%.
Fig. 8.7 is a plot of Ca/Mg versus depth, and there appears to be an
increase of this ratio with depth which would indicate that magnesium is
depleted more with respect to calcium in brines taken from deeper strata.

7 .• . .•
6

5
. .•
°1'"
u::;: •
4 •
3
••
...

2 ~
1
20 80 140 200 260 320
DISSOLVEO SOllDS, g/I

Fig. 8.4. Plot of the Ca/Mg ratio versus dissolved solids for some formation waters taken
from the Rodessa forrnation in the East Texas Basin. The line is a least squares fit
2
(y = a + bx + cx ) for the scattered data.

70
~
o Cretaceaus gulf
60-
.• Cretaceous camanche

f--
ó Jurassic
50
x Pennsylvanian
o
• Tertiary
40 f--

°l~
Z +
u
o 30

20
-

-
.o

.o o o o

..o,
o
10
1 ti o o
o • o
.. ~, x
.o
~ .o
fX ,;¡¡. o
O 1 1 ·1
500 1,000 1,500 2,000 2,500 3,000
DEPTH, meters

Fig. 8.5. Plot of the Na/Ca + Mg) ratio versus depth for some formation waters taken
from sedimentary rocks in the East Texas Basin
APPLICATION OF THE CLASSIFICATION SYSTEMS 287

8.-._--------------------------------------,

0>5
o::.

o
1u
4

3
..
2

L- ~~----~-----L~.--~~I __----~I-----
lO 500 1,000 1,500 2,000 2,500 ·3,000
DEPTH, rnet ers

Fig. 8.6. Plot of the Na/(Ca + Mg) ratio versus depth for some formation waters taken
from the Rodessa formation in the East Texas Basin. The line is a least squares fit
(y = a + bx + ex 2) for the scattered data.

KEY
.
30- o Cretoceous gulf
• Cretoceous comonche
II Jurossic
x Pennsylvonion
o • Tertiory o
20-

• o
l$>
II
• II

10 - • o
• !' o
8 o

-g
S
8
o
o.~

•I
• •
•••

, ••
&

. o

~ •• x
• ..
O I • I I
I ••

500 1,000 1,500 2,000 2,500 3,000


DEPTH, meters

Fig. 8.7. Plot of the Ca/Mg ratio versus depth for so me formation waters taken from
sedimentary rocks in the East Texas Basin.

This would indicate that magnesium is taken from the brine by reactions to
form minerals or absorbed by clays.
Fig. 8.8 is a plot of the least squares analysis data for Ca/Mg versus depth
for some brines from the Rodessa formation in the East Texas Basin. The
plot indicates that the concentration of calcium increases with depth with
288 CLASSIFICATION OF OILFIELD WATERS

8.------------------------------------------,

5
.- .
01'"
ü;:< ..
4

2 ..,
IL- -L L- ~ ~ ~L_ -U

O 500 1,000 1,500 2.000 2,500 3.000


DEPT H, meters

Fig. 8.8. Plot of the Ca/Mg ratio versus depth for some formation waters taken from the
Rodessa forma tion in the East Texas Basin. The line is a least squares fit (y = a
+ bx + ex 2 ) for the scat tered data.

respect to magnesium. Reactions occurred between the brines and the


enclosing rocks to deplete the magnesium or increase the calcium in these
brines. The curve in Fig. 8.8 represents a 35% mean fit of the scattered data
where an ideal fit is 100%.

Ternary diagram

Temary or triangular diagrams are useful in studying the distribution of


the cations and anions in subsurface waters. The ions which often are used in
these diagrams are sodium, calcium, and magnesium or chloride, sulfate, and
bicarbonate. To plot these constituents, the equivalent weights of three
cation constituents or three anion constituents are determined and summed
to equal 100%. The percentage of the three then are plotted. Dickey (1966)
used these types of plotsto study the composition of some deep subsurface
waters.
Fig. 8.9 is a ternary plot of the normalized values for sodium, calcium,
and magnesium concentrations of some brines taken from oil-productive
sedimentary basins. This diagram indicates that all of these brines contain
more calcium than magnesium. For these particular samples the diagram
seems to indicate that the calcium equivalents tend to increase with the age
of the rocks fram which the brines were taken; this do es not, however, apply
to all brines.
DISCUSSION 289

No

• Terliory
O uccer cr etcceous
• Lower cr etoceou s
Ó Jurossic
X Pennsylvonion


x
X

CoL-----------------------------------------------------------~M9
Fig. 8.9. Relative concentrations of sodium, calcium, and magnesium in some formation
waters taken from sedimentary rocks in oil-productive basins.

Discussion

Chebotarev (1955) postulated that most subsurface waters are altered by


meteoric water. The three hydrodynamic zones that he considers are: (1)
active exchange, where water is influenced by a relatively high degree of
hydrodynamic movement with flushing-out of most of the salt thus pro-
ducing a low salinity water; (2) delayed exchange, where the hydrodynamic
flow is less rapid, thus leaving a higher salinity water; and (3) stagnant
conditions, where there is little hydrodynamic movement, so that the
salinity accumulates and is higher.
This system was not applied to the 4,000 waters because it evaluates in a
rather complex manner the types and classes of waters from the constituent
compositions in reacting values in percent. The method appears useful in
certain general types of studies.
Palmer's system does not consider ionic concentrations or possible con-
ditions of saturation of calcium carbonate or calcium sulfate. Some consti-
tuents such as the alkaline earths and chloride and sulfate are lumped togeth-
290 CLASSIFICATION OF OILFIELD WATERS

er. Calcium and magnesium should not be lumped together because even
though they are in the same chemical group in the Periodic Table, they often
behave quite differently in chemical reactions. For example, the solubilities
of many of their salts are considerably different. The same often is true of
the chlorides and sulfates.
The portion of the Palmer system applied to the 4,000 brines was used
only because it is incorporated in Sulin's system. The Palmer system by itself
appears to have little value in studying subsurface brin es related to hydrocar-
bon accumulations.
Schoeller separates the waters into six types solely on the basis of the
concentration of the chloride ion and into four additional types using the
concentration of the sulfate ion. Because of this, his classification of oilfield
waters gives considerably more variation in types than the Sulin system, and
thus it is more confusing in interpretation. Schoeller's formula for deter-
mining relative saturations of sulfates and carbonates in the waters has value
in production problems. For example, knowledge that the sulfates or carbon-
ates may precipitate fram a water under certain conditions is useful so that
correct treatment can be applied to prevent well and/or formation damage.
The index of base exchange also is useful in evaluating diagenetic reactions
that may have occurred to change the water.
Sulin's system considers some of the ion s in determining the type of
water, and the class indicates the predominance of the anion groups. These
two characteristics of water s appear useful in studying waters that are likely
to be associated with hydrocarbon accumulations. For example, several in-
vestigators have determined that the chloride-calcium type in the SI S2 A2
class of water is the most likely type to be associated with a hydrostatic
environment that pro motes the accumulation of hydrocarbons. The next
most likely type is bicarbonate-sodium in the SI Al A2 class and the least
likely types are chloride-magnesium and sulfate-sodium.
Knowledge of the type and class plus what Sulin describes as the signifi-
cant indicators (direct and indirect) appears useful in hydracarbon explora-
tion studies. The Sulin system does not consider the degree of saturation of
the water with respect to sulfates and carbonates, which is a disadvantage in
production problems. Also, the system makes no provision for determining
the degree of base exchange which is useful in certain interpretations of
diagenesis.
Bojarski's modification of the Sulin system appears to have little value in
evaluating a water that is likely to be associated with petroleum. Many of the
values that were found for the Na/CI ratio in the 4,000 samples were greater
than 0.85. This value in a chloride-calcium type water according to Bojarski
is not likely to be found in a water associated with petraleum. It is possible
that the waters that Bojarski evaluated were consistently very concentrated
waters with respect to dissolved solids.
CONCLUSIONS 291

Conclusions

Classification of the subsurface waters aids in interpreting the type of


water that is more likely to be associated with a hydrocarbon accumulation.
Application of water classification techniques to 4,000 oilfield brine samples
revealed that the majority of the samples were chloride-calcium type of the
81 82 A2 class, had positive IBE values, had a predominant cation sequence
of Na > Ca > Mg, and were very highly concentrated with dissolved chloride.
Many of the 4,000 samples had values greater than 7 for .v(HC03 +
C03 )2 (Ca), and some had values greater than 70 for ylCa x 804, indicating
that many of the waters were saturated or nearly saturated with respect to
carbonate and sulfate.
The Bojarski modification of the Sulin system appears to be of little value.
The ratio Naj(Ca + Mg) generally decreased with increasing dissolved solids
and depth. The ratio Ca/Mg generally increased with increasing dissolved
solids and depth.
8tudies of the water characteristics and mapping of the more important
water properties in conjunctíon with other geological and geophysical data
appear useful in exploration. Maps of certain water characteristics such as
the JCa+2 x 804 2 and .y(HC03 + C03 -2)2 (Ca+2) are useful in solving
production problems.

References

Ambrose, A.W., 1921. Underground conditions in oilfields. US. Bur. Min. Bull., 195, 238
pp,
American Petroleum Institute, 1968. API Recommended Practice [or Analysis of Oilfield
waters. Subcommittee on Analysis of Oilfield Waters, API RP 45, 2nd ed., 49 pp.
Bojarski, L., 1970. Die Anwendung der hydrochemischen Klassifikation bei Sucharbeiten
auf Erdül. Z. Angew. Geol., 16:123-125.
Chebotarev, 1.1., 1955. Metamorphism of natural waters in the crust of weathering.
Geochim. Cosmochim. Acta, 8:22-48, 137-170, 198-212.
Collom, R.E., 1922. Prospecting and testing for oil and gas. US. Bur. Min. Bull., 201, 170
pp.
Dickey, P.A., 1966. Patterns of chemical composition in deep surface waters. Bull. Am.
Assoc. Pet. Geol., 50:2472-2478.
Elliott, Jr., W.C., 1953. Chemical characteristics of waters from Canyon, Strawn, and
Wolfcamp formations in Scurry, Kent, Borden, and Howard Counties, Texas. Peto
Eng., 25 :B77-B89.
Mills, R. van A., 1925. Protection of oil and gas field equipment against corrosion. US.
Bur. Min. Bull., 233, 127 pp.
Ostroff, A.G., 1967. Comparison of some formation water classification systems. Bul/.
Am. Assoc. Pet. Geol., 51 :404-416.
Palmer, C., 1911. The geochemical interpretation of water analyses. U.S. Geol. Surv.
Bull., 749: 5-31.
Reistle, Jr., C.E., 1927. Identification of oilfield waters by chemical analysis. US. Bur.
Min. Tech. Paper, No.404, 24 pp.
Reistle, Jr., C.E. and Lane, E.C., 1928. A system of analysis of oilfield waters. US. Bur.
Min. Tech. Paper, No.432, 14 pp.
292 CLASSIFICATION OF OILFIELD WATERS

Rogers, G.S., 1917. Chemical relations of the oilfield waters in San Joaquin Valley,
California. U.S. Geol. Surv. Bull., 653:5-116.
Rogers, G.S., 1919. Sunset-Midway oil field, California, 11. Geochemical relations of oil,
gas, and water. U.S. Geol. Surv. Pro]. Paper, No.117, 103 pp.
Schoeller, H., 1955. Geochemie des eaux souterraines. Rev. Inst. Fr. Pet., 10:181-213,
219-246, 507-552.
Sulin, V.A., 1946. Waters of Petroleum Formations in the System of Nature Waters.
Gostoptekhizdát, Moscow, 96 pp. (in Russian).
Swigart, T.E. and Schwarzenbek, F.X., 1921. Petroleum engineering in the Hewitt
Oilfield, Oklahoma. Cooperative Bulletin published by the U.S. Bureau of Mines, the
State of Oklahoma, and the Ardmore, Oklahoma, Chamber of Commerce, 61 pp.
U.S. Bureau of Mines, 1965. Analyses of Oilfield Waters. Open-file Report.

You might also like