Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

The A m e r i c a n Association iif Pflroleuni (icologists Bullclin

V. 7 l , N o . 6 ( J i i n L ' 1987). F' 633-642, 6 higs.

Assessing the Relative Importance of


Compaction Processes and Cementation to
Reduction of Porosity in Sandstones^
DAVID W. HOUSEKNECHT'

ABSTRACT INTRODUCTION

At the depositional surface, well-sorted sand has Sandstone reservoir quality is largely determined by dia-
approximately 40% porosity. During burial diagenesis, genetic processes that either reduce or enhance porosity
that porosity is reduced by mechanical compaction, inter- Mechanical compaction, intergranular pressure solution,
granular pressure solution, and cementation. Mechanical cementation, framework grain dissolution, and cement dis-
compaction and intergranular pressure solution can both solution have all been documented as playing significant
be considered compactional processes because they irre- roles in modifying porosity of various sandstones. Never-
versibly reduce the intergranular volume of sand. In con- theless, current literature tends to emphasize that cementa-
trast, cementation occludes, but does not reduce, tion reduces porosity and that framework grain and cement
intergranular volume. dissolution enhance porosity, and tends to ignore the impor-
The relative importance of compactional processes and tance of mechanical compaction and intergranular pressure
cementation to porosity reduction can be quantified using solution. For example, AAPG's recently published memoir
a graph of intergranular volume vs. cement. This diagram on clastic diagenesis (McDonald and Surdam, 1984) con-
can be used to evaluate which diagenetic processes have tains 22 papers, of which only four briefly mention the role
been most influential to intergranular porosity reduction of compactional processes in determining ultimate reservoir
and to determine why some sandstones retain better reser- quality.
voir quality than others. The diagram can also be used to In contrast to the other diagenetic processes, mechanical
reconstruct pathways taken by sandstones during burial compaction and intergranular pressure solution reduce the
diagenesis. intergranular volume (minus cement porosity) of sands and
Results of applying this technique to data from the Nug- sandstones. The intergranular volume eliminated by these
get Sandstone and Bromide sandstone (Simpson Group) processes might otherwise have been occupied by pore
indicate that mechanical compaction and intergranular space or by some soluble cement (e.g., calcite) that could
pressure solution were much more important than cemen- potentially dissolve during later diagenesis. Thus, reducing
tation in determining ultimate porosity. Moreover, the best the intergranular volume by these two processes represents
porosity is preserved in samples that have undergone the irreversible destruction of porosity and potential porosity in
least intergranular pressure solution. These conclusions sandstones.
emphasize the importance of integrating an evaluation of This paper presents a technique with which the relative
these compactional processes into analyses of reservoir importance of compactional processes and cementation to
sandstones and into models of burial diagenesis. porosity reduction can be quantified, and illustrates the use
of the technique with data from two reservoir sandstones.
This technique is intended for use on sandstones composed
primarily of nonductile grains and is useful for analyzing
©Copyright 1987. The American Association of Petroleum Geologists. All
rights reserved.
intergranular porosity only; it cannot be used to analyze
^Manuscript received, June 16,1986; accepted, January 19,1987. secondary porosity generated by dissolution of framework
^Department of Geology University of Missouri, Columbia, Missouri 65211. grains.
Acl<nowledgment is made to the donors of The Petroleum Research Fund,
administered by the American Chemical Society, for partial support of this
research (14417-AC2). Amoco Production Company (for work on the Nugget
Sandstone) and Tenneco Oil Company (for work on the Simpson Group Bro-
mide sandstone) also provided partial support. SEM work was performed in the ORIGINAL POROSITY OF SAND
University of Missouri Geology SEM facility, managed by Lou Ross. The
National Science Foundation (EAR-8217931); Amoco Foundation, Inc.; Chev-
ron U.S.A.; Marathon Oil Company; Mobii Oil Corporation; Phillips Petroleum The porosity of naturally deposited sand is difficult to
Company; and Texaco Philanthropic Foundation, Inc., provided funding for
the SEM facility. The cathodoluminescence microscope used in this research measure because sampling procedures tend to disturb origi-
was partially funded by the Research Council of the Graduate School, Univer- nal packing arrangements. The most extensive data base
sity of Missouri-Columbia. collected from natural sands is that of Pryor (1973), who
I sincerely thank my friends at Amoco, Chevron, and Exxon research labs for
numerous discussions and suggestions that helped mold my perceptions of measured the porosity, permeability, and textural character-
sandstone diagenesis. Friends at Exxon convinced me that "intergranular vol- istics of sands from river point bars, beaches, and eolian
ume" is more appropriately descriptive than terms I used in the past. Finally, I
thank Earle McBride, Lee Stephenson, and Donna Summers for thoughtful
dunes adjacent to beaches. Porosity values were found to
reviews that improved this paper. vary considerably within deposits of each environment.

633
634 Compaction Processes and Cementation

Point bar sand porosity ranged from 17 to 52%, with a REDUCTION OF POROSITY BY
mean of 41 %; beach sand porosity ranged from 39 to 56%, COMPACTION AND CEMENTATION
with a mean of 49%; and dune sand porosity ranged from
Operational Definitions
42 to 55%, with a mean of 49%. However, most of the low
porosity values in point bar sands represents poorly to mod-
erately sorted samples; in well and very well sorted samples, Three diagenetic processes are important in modifying
no porosity values below 40% were reported (Pryor, 1973, intergranular porosity: mechanical compaction, chemical
his Figure 18). Beard and Weyl (1973) determined that the compaction, and cementation. Before the relative impor-
porosity of 25 Holocene barrier-island sand samples ranged tance of each can be quantified, clear definitions of these
from 40 to 45%. Others have reported porosity values to processes must be estabUshed.
range from 35 to 45% in sands from a variety of modern Mechanical compaction. —Mechanical compaction is the
environments (e.g., Eraser, 1935; Hamilton and Menard, bulk volume reduction resulting from processes other than
1956). framework grain dissolution. Generally induced by litho-
Numerous researchers investigated the relationships static stress, mechanical compaction is characterized by the
among texture, porosity, and permeability of unconsoli- reorientation and repacking of competent (brittle) grains,
dated, natural sand by making "artificial" grain packs in by local fracture or cleavage of brittle grams, and by plastic
the laboratory. The most comprehensive study is that of deformation of ductile grains. In sands composed predomi-
Beard and Weyl (1973). After mixmg sand into discrete sort- nately of nonductile grains, mechanical compaction does
ing categories and artificially packing the sand in water, not reduce the volume of framework grains; mechanical
they reported mean poroshies as follows: very poorly compaction is characterized entirely by the reduction of
sorted, 27.9%; poorly sorted, 30.7%; moderately sorted, intergranular volume and, therefore, porosity
34.0%; well sorted, 39.0%; very well sorted, 40.8%; and Chemical compaction.—Chemicsl compaction is the
extremely well sorted, 42.4%. Beard and Weyl (1973) also bulk volume reduction caused by the dissolution of frame-
artificially packed sand samples collected from Holocene work grains at points of contact (also see Fiichtbauer, 1967).
barrier islands and found the sand to be well to extremely Chemical compaction is generally induced by lithostatic
well sorted and the porosities to range from 38 to 43%. stress and is characterized by intergranular pressure solu-
They found excellent agreement (within 3% porosity) tion, although the definition can include stylolitization that
among the porosities of naturally packed sand, artificially may be induced by either lithostatic or tectonic stress.
packed sand with natural sorting, and artificially packed Chemical compaction reduces the volume of framework
sand with sorting controlled by laboratory mixing of sieved grains; it also reduces intergranular volume and, therefore,
sand fractions. In similar, but less comprehensive studies, porosity by causing the closer packing of framework grains
numerous other researchers found porosities to range from (e.g., Rittenhouse, 1971a).
35 to 47% for sands well sorted and better (e.g., Gaither, Cementation.—Csmsntation is the occlusion of inter-
1953; Rogers and Head, 1961). granular volume by the precipitation of authigenic miner-
An alternative approach to determining relationships als, with no directly related reduction of bulk volume.
among texture, porosity, and permeability has been to work Cementation always results in the reduction of intergranu-
with geometric packing of perfect spheres. The landmark lar porosity.
work of Fraser (1935) and Graton and Fraser (1935) pro-
vided the basic concepts of how grain size, sorting, and Conceptualization
grain packing affect porosity. Fraser and Graton demon-
strated that porosity is independent of grain size, but When assessing the diagenetic modification of intergran-
decreases significantly with poorer sorting. They also estab- ular porosity, it is useful to separate the effects of compac-
lished the existence of a limited number of ideal packing tional processes from the effects of cementation by using
arrangements for perfectly sorted spheres and calculated the diagram shown in Figure 1. The vertical axis represents
the porosity associated with each, ranging from a minimum intergranular volume (expressed as a percentage of the
of 25.95% for rhombohedral (tightest) packing to 47.64% whole rock), which is the sum of intergranular porosity plus
for cubic Ooosest) packing. These ideal packing models pro- all cements that occupy intergranular space. The term
vide a conceptual framework within which one may visual- "intergranular volume" was apparently first used by Weller
ize natural intergranular porosity. (1959), and is synonymous with minus-cement porosity
The research cited above suggests that most sands that (Rosenfeld, 1949; Heald, 1956). This variable is easily
are well sorted or better contain at least 40% porosity at the quantified by point counting, although the point counting
depositional surface. That value (40%) is used throughout must be performed on a cathodoluminescence microscope
this paper as the original porosity that is subsequently modi- if, on a standard petrographic microscope, the outlines of
fied by diagenetic processes during burial. Variations in detrital grains are not clearly distinguishable from cements
original porosity will affect the results obtained from the (e.g., Houseknecht, 1984).
technique discussed below, and these variations will be Compactional processes determine the modification of
addressed in a subsequent section. This technique analyzes intergranular volume during burial, as outlined in the pre-
the modification of intergranular porosity only (including vious section. At the depositional surface, a well-sorted
cement-dissolution porosity), and does not include the sand has an intergranular volume (i.e., intergranular poros-
development of grain-dissolution porosity ity, inasmuch as there is no cement present) of about 40%;
David W. Houseknecht 635

CEMENT / INTERGRANULAR VOLUME fragments can undergo a total destruction of intergranular


0 0.5 1.0 volume during mechanical compaction as a result of plastic
deformation of the lithic fragments and the generation of
pseudomatrix (Dickinson, 1970; Rittenhouse 1971b). This
paper deals only with sandstones composed predominately
of nonductile framework grains; the presence of abundant
ductile lithic fragments has not yet been incorporated into
the analysis.
The extent to which chemical compaction can reduce the
intergranular volume of nonductile-grain sands depends on
one's perspective of mechanical compaction. If we assume
that mechanical compaction can reduce intergranular vol-
ume to about 26%, then any further reduction can occur
only through chemical compaction. Alternatively, if we
assume that mechanical compaction does not significantly
reduce intergranular volume, then any reduction of inter-
granular volume below 40% can be attributed to chemical
compaction. Heald (1956) demonstrated that intergranular
pressure solution can destroy virtually all intergranular vol-
ume in certain situations.
The horizontal axis of Figure 1 represents cementation;
the proportion of intergranular volume occluded by cement
is the quantity plotted on this axis. (Again, this variable is
Figure 1—Diagram Ulustrating how intergranular porosity of calculated from point-count data easily collected during
sandstones is dependent upon intergranular volume and propor-
tion of intergranular volume that is occluded by cement. Vertical routine petrographic analysis.)
axis represents compactional processes, horizontal axis repre- As shown by the curved lines on Figure 1, the volume of
sents cementation. A, B, C, and D are data plotted for end- intergranular porosity present in a sandstone is a function
member sandstones discussed in text and illustrated in Figures 2 of how much intergranular volume has been destroyed by
and 3. mechanical and chemical compaction (vertical axis) and
how much of that intergranular volume is occluded by
cement (horizontal axis). For example, a sandstone that
during burial, that sjmd may undergo mechanical compac- contains 10% intergranular porosity may plot anywhere
tion and chemical compaction, both of which will reduce its along the 10% curve of Figure 1. The sandstone's porosity
intergranular volume. The relative importance of these two may be predominately a function of mechanical and chemi-
compactional processes is difficult to establish and is, there- cal compaction (in which case it would plot close to the ver-
fore, rather controversial. It is widely assumed that, in a tical axis), may be predominately a function of cementation
sand composed of nonductile grains that are fairly spherical (in which case it would plot close to the horizontal axis), or
and well rounded, intergranular volume potentially can be may be a function of both compaction and cementation (in
reduced to 26% (approximate closest packing) by mechani- which case it would plot at some point between the two
cal compaction alone, a 35% reduction in original inter- axes).
granular volume. However, some argue that it is impossible Note that many previous workers have attempted to esti-
to reduce the intergranular volume of a sand below about mate the effects of compactional processes by quantifying
40% by pure mechanical compaction, without introducing grain packing in various ways. Noteworthy are the works of
a source of negative entropy (Lee Stephenson, personal Kahn (1956a, b) and the summary provided by Griffiths
communication). Essentially, the intergranular volume (1967). None of these proposed techniques has gained wide-
would have to be increased in some way and then the grains spread popularity, probably because they are difficult to
repacked by some process that is more efficient than the apply in a quantitative way.
original depositional medium. Finally, other workers
argued that mechanical and chemical compaction work
together to reduce the intergranular volume of sand during Petrographic Characteristics
shallow burial. For example, Fiichtbauer (1967) demon-
strated how small amounts of intergranular pressure solu- Figures 2 and 3 illustrate the petrographic characteristics
tion can remove corners of grains and allow them to slide of sandstones that occupy end-member positions on the
past each other to assume a tighter packing arrangement. compaction-cementation diagram of Figure 1. In Figures 2
Thus, depending on one's point of view, mechanical com- and 3, the letter with which each micrograph is labeled cor-
paction may be a significant process that reduces intergran- responds to the letters shown on Figure 1. Sandstones origi-
ular volume to as Uttle as 26% or may be a relatively nally composed almost exclusively of spherical and well
insignificant process that hardly modifies the original inter- rounded, monocrystalline quartz grains have been chosen
granular volume established by the depositional medium. for these illustrations because they represent the simplest
Sands containing significant volumes of ductile lithic and most easily visuahzed situations possible.
636 Compaction Processes and Cementation
David W. Houseknecht 637

Figure 2—Thin-section micrographs of sandstones that display end-member compaction-cementation characteristics. (A) Plane-
polarized light photomicrograph of sandstone containing 30% intergranular volume and virtually no cement. All grains are quartz,
and pores are filled with blue stained epoxy. St. Peter Sandstone, St. Qair County, Illinois; depth = 2,164 ft. (B) Cathodolumines-
cence micrograph of sandstone containing 29'fo intergranular volume, most of which is occluded by quartz cement (overgrowths).
Quartz grains luminesce various shades of red and blue; quartz cement is nonluminescent to slightly luminescent in shades of red;
porosity appears black. Bromide sandstone, McClain County, Oklahoma; depth = 10,184 ft. (C) Cathodoluminescence micrograph
of sandstone containing 5<Vo intergranular volume and virtually no cement. Luminescence is the same as described for (B). Bromide
sandstone, McClain County, Oklahoma; depth := 10,175 ft. (D) Cathodoluminescence micrograph of sandstone containing 6% inter-
granular volume, most of which is occluded by quartz cement (overgrowths). Luminescence is the same as described for (B), except
that here, black areas represent nonluminescent illite, which is easUy observed in plane light. This sample contains virtually no poros-
ity. Bromide sandstone, Cleveland County, Oklahoma; depth = 8,132 ft. All scale bars = 0.25 mm.

Figures 2A and 3 A illustrate typical thin section and scan- low value required significant chemical compaction. The
ning electron microscope (SEM) views of a sandstone that modification of grain shape and the nature of grain con-
has undergone a minor reduction of intergranular volume tacts confirms this inference. Carbonate cemented nodules
and virtually no cementation. Intergranular volume in this within this sandstone display large intergranular volumes
sandstone is 30%, as determined by point counting (300 and well-sorted quartz grains that are as spherical and well
points). Following the line of reasoning presented earlier, rounded as those shown in Figures 2A and 2B (Bajsarowicz
we might expect that this sandstone has undergone some and Houseknecht, 1983). The nonspherical and poorly
mechanical compaction, but probably no chemical com- rounded appearance of most of the quartz grains in Figure
paction. Thin-section analysis confirms this because most 2C can thus be attributed to shape modification by inter-
of the quartz grains display either no contacts (floating granular pressure solution that occurred throughout most
grains) or tangential contacts with neighboring grains (Fig- of the sandstone where carbonate cement is absent. Addi-
ure 2A). Very few long and concave-convex contacts are vis- tionally, the predominance of concave-convex and sutured
ible. However, the scanning electron micrograph of Figure contacts and the absence of floating grains and tangential
3A clearly shows the presence of small pressure solution contacts indicate that a significant amount of intergranular
scars, indicating that at least slight intergranular pressure pressure solution has occurred. SEM examination of sand-
solution has occurred. This observation is significant stones that have undergone so much chemical compaction
because even small increments of intergranular pressure is difficult because samples tend to break through, rather
solution significantly reduce intergranular volume (Rit- than around, grains during sample preparation. Neverthe-
tenhouse, 1971a). Thus, it appears that mechanical com- less, the predominance of grain contacts indicative of pres-
paction does not necessarily reduce intergranular volume to sure solution and the small intergranular porosity are
the minimum possible value (26%) before chemical com- diagnostic of a sandstone that has undergone extreme
paction begins. For this reason, it is difficult to assess pre- chemical compaction (Figure 3C). Sandstones that have
cisely how much of the reduction in intergranular volume such low intergranular volumes also have very low permea-
results from mechanical compaction and how much results bilities because the remaining porosity tends to be isolated
from chemical compaction, even in sandstones that have and, therefore, noneffective.
large intergranular volumes preserved. Figures 2D and 3D illustrate typical thin section and SEM
Figures 2B and 3B illustrate typical thin section and SEM views of a sandstone that has undergone a major reduction
views of a sandstone that has undergone a minor reduction of intergranular volume and quartz cementation. Inter-
of intergranular volume and significant quartz cementa- granular volume in this sandstone is 6%, so we can assume
tion. Intergranular volume in this sandstone is 29%, so we that its compaction history is similar to that of the sand-
can assume that its compaction history is similar to that of stone discussed in the preceding paragraph. Grain-shape
the sandstone discussed in the preceding paragraph. In modification and the geometry of grain contacts are identi-
cathodoluminescence (Figure 2B), the grain contact geome- cal to those shown in Figure 2C. However, in this sandstone
tries are similar to those shown in Figure 2A. This sample all of the intergranular volume is occluded by quartz and ill-
emphasizes the importance of using cathodoluminescence; ite cements, so there is essentially no intergranular porosity
neither the intergranular volume nor the nature of grain preserved.
contacts could be accurately determined using standard pet-
rographic techniques or scanning electron microscopy
because of the abundance of quartz cement. Recognizing
QUANTinCATION
samples that fall into the upper-right portion of Figure 1 is
particularly important in dealing with sandstones that con-
tain relatively soluble cements, such as calcite. These Figure 1 is useful for conceptualizing the relative roles of
cements may allow large volumes of cement-dissolution compaction and cementation in modifying intergranular
porosity to be generated if suitable geochemical conditions porosity, but it is somewhat awkward to use for analyzing
occur. large volumes of data. As an alternative, Figure 4 can be
Figures 2C and 3C illustrate typical thin section and SEM used to rapidly and accurately assess the relative importance
views of a sandstone that has undergone a major reduction of compaction and cementation, and to calculate intergran-
of intergranular volume and virtually no cementation. ular porosity values from point-count data. The vertical
Intergranular volume in this sandstone is 5%, and no axis of Figure 4 represents intergranular volume, with the
cement is present. Reducing intergranular volume to such a assumption that well-sorted sand has an original porosity of
638 Compaction Processes and Cementation
David W. Houseknecht 639

Figure 3—Scanning electron micrographs of Figure 2 sandstones, displaying end-member compaction-cementation characteristics.
(A) St. Peter sample displays large intergranular volume, very little quartz cement, and small grain-contact scars (dimples) caused by
intergranular pressure solution. (B) Bromide sample displays large intergranular volume that is almost totally occluded by quartz
cement. Note relatively small grain-contact scars (ovoid, rough-looking areas). (C) Bromide sample displays small intergranular vol-
ume, very little quartz cement, and sutured grain-contact scars that are large and pervasive. (D) Bromide sample displays small inter-
granular volume, most of which is occluded by quartz cement (e.g., lower left), and sutured grain-contact scars that are large and
pervasive. All scale bars = 0.10 mm.

40%. As shown on Figure 4, this axis can also be used to The dashed, diagonal line on Figure 4 separates samples
quantify the percentage of original porosity that has been in which compaction has been more important than cemen-
destroyed by mechanical and chemical compaction; this tation (lower left) from samples in which cementation has
value can be quantified using the following equation: been more important than compaction in determining inter-
granular porosity (upper right). Thus, when individual sam-
Percent of original porosity destroyed by compaction ples are plotted on this diagram, a rapid assessment can be
_ 40 - intergranular volume made as to the relative importance of compaction and
X 100, (1) cementation in determining ultimate reservoir quality. The
40
where intergranular volume is expressed as a percentage of four end-member sandstones discussed previously are plot-
whole rock. ted on Figure 4 to illustrate this method. Moreover, diagen-
The horizontal axis of Figure 4 represents the percentage etic pathways can be drawn on this diagram by plotting
of cement present, and it can be used to plot the percentage intergranular volumes and cement percentages for discrete
of original porosity destroyed by cementation, with that paragenetic events based on cement-stratigraphy relation-
value quantified using the following equation: ships (an example is illustrated in a subsequent section).

Percent of original porosity destroyed by cementation


cement EXAMPLES
40
X 100, (2)

where cement is the volume of cement present expressed as a Data from two reservoir sandstones, presented in this
percentage of whole rock. section, illustrate how the techniques presented above can
Lines of equal intergranular porosity plot as straight be used to evaluate the relative importance of compaction
diagonals on this diagram, and intergranular porosity can and cementation to the development of reservoir quality.
be estimated either directly from the figure, or by using the Each example is based on detailed cathodoluminescence
following equation: petrography (3(X) points counted per sample) and scaiuiing
electron microscopy of quartzose sandstones. In each case,
Intergranular porosity = intergranular volume - cement. (3) thin sections that display a unimodal grain-size distribution
are treated as individual samples, and thin sections that dis-
play pronounced laminations defined by significant grain-
CEMENT (%)
size variation were divided so that each lamination was
treated as an individual sample during point counting (see
Houseknecht, in press, for additional details). Data were
collected following the techniques desaibed by House-
knecht (1984).

Example 1: Nugget Sandstone

The Jurassic Nugget Sandstone in the Anschutz Ranch


East field (Summit County, Utah) is fine to medium grained
and moderately to well sorted. On the average, monocrys-
talline quartz comprises 87% and feldspar comprises 11%
of the framework grains. Porosity development has been
influenced by both textural characteristics imposed by flu-
vial and eolian depositional environments and by diagenetic
processes that included mechanical compaction, intergran-
ular pressure solution, precipitation of cements, and partial
ORIGINAL POROSITY DESTROYED dissolution of cements (Lindquist, 1982; Aucremann, 1984;
BY CEMENTATION (%) Aucremann and Houseknecht, 1984). Cements present
include ilUte, dolomite, calcite, and quartz, although none
Figure 4—Diagram for evaluating relative importance of com-
pactional processes and cementation to porosity development. of these are pervasively abundant. At least some porosity
A, B, C, and D are data plotted for end-member sandstones dis- enhancement, caused by the dissolution of carbonate
cussed in text and illustrated in Figures 2 and 3. cements, seems to have occurred.
640 Compaction Processes and Cementation

CEMENT (%) CEMENT (%)

ORIGINAL POROSITY DESTROYED BY CEMENTATION (%) ORIGINAL POROSITY DESTROYED BY CEMENTATION (%)

Figure 5—Intergranular volume vs. cement data for Nugget Figure 6—Intergranular volume vs. cement data for Bromide
Sandstone, Anschutz Ranch East field, Utah. Dots represent sandstone, central Oklahoma platform. Dots represent well-
well-sorted samples, circles represent moderately sorted samples, sorted samples, and circles represent moderately sorted samples.
and star represents mean of all samples. Total number of samples Squares represent samples that contain both quartz and caldte
plotted = 111. cement, with empty squares indicating quartz only, and squares
enclosing dots and circles indicating total cement (quartz plus
calcite). Star represents mean of all samples. Total number of
The intergranular volume-cement diagram in Figure 5 samples plotted = 69. Arrows A, B, and C represent diagenetic
shows Nugget data. Note that 97% of the data points clus- pathways, as explained in text.
ter in the lower-left portion of the diagram, indicating that
these samples have undergone significant reduction of plot close to the vertical axis of Figure 5, indicating that
intergranular volume by compactional processes and rela- compactional processes were more important than cemen-
tively little cementation. Sorting appears to have played an tation in determining ultimate porosity. Moreover, the high-
insignificant role in determining intergranular volume; est porosity samples contain about 10% more intergranular
moderately and well-sorted samples display no preferred volume and about 5% less cement than average. Thus, the
distribution on the diagram. This distribution suggests that best quality reservoir samples are those that have undergone
using a single original value of intergranular volume (40%, less destruction of intergranular volume by intergranular
in this case) is a reasonable assumption. The wide range in pressure solution, and that contain less cement than aver-
intergranular-volume values is the result of varied amounts age, although the former is somewhat more important.
of intergranular pressure solution, which was largely con-
trolled by the variation in grain size and clay content, as
explained by Houseknecht (in press). The mean values for
Example 2: Simpson-Bromide Sandstone
all data (indicated by the star on Figure 5) are 15.3% inter-
granular volume and 10.8% cement (which includes all
cements). Assuming that the average original porosity of The Ordovician Bromide sandstone (Simpson Group of
the Nugget was 40%, then 62% of the original porosity has the central Oklahoma platform) is fine to medium grained
been destroyed by compactional processes, whereas only and moderately to very well sorted. On the average, 99% of
27% has been destroyed by cementation. Petrographic the framework grains are monocrystalline quartz. Porosity
observations support these results, as most grains display development has been influenced by mechanical compac-
concave-convex or sutured contacts, or both, as well as tion, intergranular pressure solution, and the precipitation
other evidence of significant intergranular pressure solution of cements that include quartz, illite, and local calcite
(Aucremann, 1984; Houseknecht, in press). (Heald, 1956; Bajsarowicz, 1983; Bajsarowicz and
Note that most of this study's samples were collected Houseknecht, 1983).
from cores of the upper and lower portions of the Nugget The intergranular volume-cement diagram of Figure 6
Sandstone, and only a few were available from the middle, shows the Bromide data. The data plotted as dots and cir-
highest porosity portion of the reservoir. Thus, most of the cles reflect the sum of all cements present. Note that 70% of
Nugget samples plot between the 0 and 10% intergranular the samples plot in the lower left portion of the diagram,
porosity lines on Figure 5, and the average porosity for all indicating that a larger percentage of their original porosity
samples, calculated using equation 3, is only 4.5%. The five has been destroyed by compactional processes than by
samples that plot between the 10 and 20% lines on Figure 5 cementation. As with the Nugget, sorting does not appear
are from the middle portion of the reservoir, where both to influence significantly the intergranular volumes of the
porosities and permeabilities are higher. Even these samples Bromide. The wide range of intergranular-volume values is
David W. Houseknecht 641

the result of varied amounts of intergranular pressure solu- ity, sorting, grain shape, and roundness. The choice of 40%
tion, which was controlled by variation in grain size and as the original porosity of unconsolidated sand may not be
clay content (Heald, 1956; Houseknecht, in press). The appropriate in all cases. For example, if a sandstone is
mean values for all data (indicated by the star on Figure 6) poorly sorted, its original porosity would certainly have
are 20.6% intergranular volume and 15.2% cement (which been less than 40%. On the other hand, if a sandstone is
includes all cements). Assuming that the original porosity extremely well sorted, composed of very angular grains, or
of the Bromide was 40%, then 48% of the original porosity displays very large values of intergranular volume within
has been destroyed by compactional processes, whereas nodules of cement precipitated during shallow burial, then
only 38% has been destroyed by cementation. The predom- its original porosity would likely have been greater than
inance of concave-convex and sutured contacts, as illus- 40%. In cases such as these, the figures and equations pre-
trated in Figures 2C, 2D, 3C, and 3D (also see House- sented herein will yield distorted estimates of the relative
knecht, in press), indicates the importance of chemical com- importance of compaction vs. cementation. This problem
paction. can be solved by substituting more appropriate values for
Each of the five samples that contain significant volumes original porosity into the figures and equations. For exam-
of calcite cement is plotted as two separate points indicated ple, if evidence suggests that the original porosity of a sand
by squares in Figure 6; empty squares include only quartz was 50%, Figures 1 and 4 can be redrawn with that value as
cement and the squares enclosing dots and circles include the maximum intergranular volume possible, and equations
the sum of quartz plus calcite. Cement-stratigraphy rela- 1 and 2 can be rewritten with 50% substituted for 40%.
tionships clearly indicate that the calcite cement postdates As yet unknown are the effects that variations in grain
the quartz cement, so the two points representing each of shape and roundness may have on results obtained using
those five samples reflect two distinct steps in the parage- these techniques. It has been shown that sands composed of
netic sequence followed by those samples. This relationship nonspherical and poorly rounded grains tend to have higher
allows generalized diagenetic pathways to be plotted in Fig- original porosities than sands composed of spherical and
ure 6. As shown by the arrows, an early phase of diagenesis well-rounded grains (Beard and Weyl, 1973). However, such
involved reducing the intergranular volume to about 30% sands tend to have a more rapid reduction of intergranular
(the mean intergranular volume of the 5 calcite-cemented volume during burial, both as the result of mechanical com-
samples is 31%). This reduction of intergranular volume paction and intergranular pressure solution (Renton et al,
probably involved both mechanical compaction and slight 1969). These relationships suggest that the compaction-
intergranular pressure solution (Bajsarowicz and cementation diagram will yield reliable estimates of the rela-
Houseknecht, 1983). At some time during or following this tive importance of compaction and cementation, but that it
compaction, an average of 12% quartz cement was precipi- may be more difficult to evaluate the relative roles of
tated. The combination of compaction and quartz cementa- mechanical compaction and intergranular pressure solution
tion suggests that the diagenetic pathway approximately in reducing intergranular volume if a sandstone is com-
followed arrow A. At that point in the diagenetic history, posed of nonspherical or poorly rounded grains, or both.
calcite cement was locally precipitated and a limited number This technique can also be applied to carbonate grain-
of samples followed arrow B. However, most samples stones composed of fairly spherical and well-rounded
underwent no calcite cementation, but were subjected to grains. For example, McLimans and Videtich (in press)
significant amounts of intergranular pressure solution dur- have demonstrated this technique's utility in analyzing
ing rapid burial to depths in excess of 3 km (10,000 ft) oolitic grainstones that contain less than 5% matrix.
(Bajsarowicz and Houseknecht, 1983). Thus, most of the Although it appears that the approach presented here can
Bromide sandstone followed arrow C, and underwent a sig- potentially be applied to a wide range of sandstones and car-
nificant reduction of intergranular volume (Houseknecht, bonate grainstones, it almost certainly will not work in
in press). sandstones that contain abundant ductile grains. The tech-
Those samples that display the best reservoir quality—they nique is also restricted to evaluating intergranular porosity
plot between the 10 and 20% intergranular porosity lines- (including cement-dissolution porosity), and is not applica-
have had less destruction of intergranular volume by inter- ble to grain-dissolution porosity.
granular pressure solution, and contain less cement than
average. These two contributing factors appear to be equally
important in the retention of relatively high porosity.
CONCLUSIONS

Intergranular porosity of sand and sandstones is deter-


DISCUSSION
mined by the amount of compaction and cementation
undergone during burial diagenesis. Compactional effects
The techniques and examples presented above demon- can be evaluated by quantifying existing intergranular vol-
strate the importance of mechanical compaction and inter- umes and by determining the nature of grain contacts.
granular pressure solution in determining the ultimate Assuming an original intergranular volume of 40% for
porosity of sandstones that are predominately composed of well-sorted sand composed of nonductile grains, that 40%
fairly spherical, well-rounded, and well-sorted grains. intergranular volume can potentially be reduced to about
Before these techniques can be indiscriminately applied, 30% by mechanical compaction (wiiich includes reorienta-
one must consider the effects of variation in original poros- tion and repacking of grains, as well as local fracture and
642 Compaction Processes and Cementation

cleavage of grains). Further reduction of intergranular vol- and D. W. Houseknecht, 1983, Relationship of paragenesis and
porosity reduction to the burial history of the Bromide sandstone,
ume can occur only as a result of chemical compaction, Simpson Group, Oklahoma (abs.): GSA Abstracts with Programs, v.
which most commonly mvolves intergranular pressure solu- 15, p. 519.
tion. This process can destroy virtually all intergranular vol- Beard, D. C , and P. K. Weyl, 1973, Influence of texture on porosity and
ume, resulting in a predominance of concave-convex and permeability of unconsolidated sand: AAPG Bulletin, v. 57, p. 349-
369.
sutured grain contacts. Porosity is also dependent upon Dickinson, W. R., 1970, Interpreting detrital modes of graywacke and
how much of the existing intergranular volume is occluded arkose: Journal of Sedimentary Petrology, v. 40, p. 695-707.
by authigenic cements. Fraser, H. J., 1935, Experimental study of the porosity and permeabiUty of
clastic sediments: Joitfnal of Geology, v. 43, p. 910-1010.
A simple graph of intergranular volume vs. cement can Fuchtbauer, H., 1967, Influence of different types of diagenesis on sand-
be used to quantify the relative roles of compactional pro- stone porosity: 7th World Petroleum Congress Proceedings, v. 2, p.
cesses and cementation in determining intergranular poros- 353-369.
Gaither, A., 1953, A study of porosity and grain relationships in experimen-
ity This diagram is useful for evaluating which diagenetic tal sands: Journalof Sedimentary Petrology, v. 23, p. 180-191.
events were most influential to porosity reduction, and to Graton, L. C , and H. J. Fraser, 1935, Systematic packing of spheres, with
determine why some sandstones retain better reservoir particular relation to porosity and permeability: Journal of Geology, v.
quality than others. Moreover, the diagram can be used to 43, p. 785-909.
Griffiths, J. C , 1967, Scientific method in analysis of sediments: New
reconstruct diagenetic pathways of reservoir sandstones by York, McGraw-HiU, 508 p.
sequentially plotting discrete paragenetic phases. Such a Hamilton, E. L., and H. W. Menard, Jr., 1956, Density and porosity of sea-
reconstruction, when coupled with time-temperature calcu- floor surface sediments off San Diego, California: AAPG Bulletin, v.
40, p. 754-761.
lations based on burial history, may be useful for estimating Heald, M. T., 1956, Cementation of Simpson and St. Peter sandstones in
reservoir quality during peak hydrocarbon generation. parts of Oklahoma, Arkansas, and Missouri: Journal of Geology, v. 64,
Data from two sandstone reservoirs demonstrate the use p. 16-30.
Houseknecht, D. W., 1984, Influence of grain size and temperature on
of the compaction-cementation diagram. The Jurassic intergranular pressure solution, quartz cementation, and porosity in a
Nugget Sandstone in the Anschutz Ranch East field shows quartzose sandstone: Journal of Sedimentary Petrology, v. 54, p. 348-
more significant reduction of intergranular porosity by 361.
compaction than by cementation. In 97% of the samples in press, Intergranular pressure solution in four quartzose sand-
stones: Journal of Sedimentary Petrology.
analyzed, mechanical compaction and intergranular pres- Kahn, J. S., 1956a, The analysis and distribution of the properties of pack-
sure solution have destroyed more porosity than has cemen- ing in sand-size sediments, 1: on the measurement of packing in sand-
tation. The Ordovician Bromide sandstone (Simpson stones: Journal of Geology, v. 64, p. 385-395.
1956b, Analysis and distribution of packing properties in sand-size
Group) frtjm the central Oklahoma platform was also sub- sediments, 2: the distribution of the packing measurements and an
jected to significant compactional processes. In 70% of the example of packing analysis: Journal of Geology, v. 64, p. 578-606.
samples analyzed, mechanical compaction and intergranu- Lindquist, S. J., 1982, Nugget Formation reservoir characteristics affecting
lar pressure solution have destroyed more porosity than has production in the overthrust belt of southwestern Wyoming: Society of
Petroleum Engineers Paper 10993.
cementation. In both of these examples, the key to predict- McDonald, D. A., and R. C. Surdam, eds., 1984, Clastic diagenesis:
ing the distribution of high porosity sandstones is under- AAPG Memoir 37,434 p.
standing the geological variables that have controlled McLimans, R. K., and P. E. Videtich, in press. Reservoir diagenesis and oil
mechanical and chemical compaction; the best quality res- migration: Middle Jurassic Great Oolite Limestone, Wealden basin,
southern England: Third Conference on Petroleum Geology of North-
ervoir rocks contain somewhat less cement than most sam- west Europe Proceedings.
ples, but also show significantly less destruction of Pryor, W. A., 1973, Permeability-porosity patterns and variations in some
intergranular volume. Holocene sand bodies: AAPG Bulletin, v. 57, p. 162-189.
Renton, J. J., M. T Heald. and C. B. Cecil, 1969, Experimental investiga-
tion of pressure solution of quartz: Journal of Sedimentary Petrology,
v. 39, p. 1107-1117.
REFERENCES CITED Rittenhouse, G., 1971 a. Pore-space reduction by solution and cementation:
AAPG BuUetin, v, 55, p. 80-91.
1971b, Mechanical compaction of sands containing different per-
AuCTemann, L. J., 1984, Diagenesis and porosity evolution of the Jurassic centages of ductile grains: a theoretical approach: AAPG Bulletin, v.
Nugget Sandstone, Anschutz Ranch East field. Summit Coimty, Utah: 55, p. 92-96.
Masterls thesis, University of Missouri, Columbia, Missouri, 116 p. Rogers, J. J., and W. B. Head, 1961, Relationships between porosity,
and D. W. Houseknecht, 1984, Carbonate cementation and disso- median size, and sorting coefficients of synthetic sands: Journal of Sed-
lution in the Nugget Sandstone, Anschutz Ranch East field, Utah imentary Petrology, v. 31, p. 467-470.
(abs.): SEPM Midyear Meeting Abstracts, p. 9-10. Rosenfeld, M. A., 1949, Some aspects of porosity and cementation: Pro-
Bajsarowicz, C. J., 1983, Diagenesis and porosity reduction of the Bromide ducers Monthly, v. 13, p. 39-42.
sandstone, Simpson Group, Oklahoma: Master's thesis. University of Weller, J. M., 1959, Compaction of sediments: AAPG Bulletin, v. 43, p.
Missouri, Columbia, Missouri, 187 p. 273-310.

You might also like