You are on page 1of 44

Earth-Science Reoiews, 26 (1989) 69-112

Elsevier Science Publishers B.V., Amsterdam--Printed in The Netherlands

69

Quartz Cement in Sandstones: A Review


EARLE F. McBRIDE

ABSTRACT McBride, E.F., 1989. Quartz cement in sandstones: a review. Earth-Sci. Rev., 26: 69-112. Quartz cement as syntaxial overgrowths is one of the two most abundant cements in sandstones. The main factors that control the amount of quartz cement in sandstones are: framework composition; residence time in the "silica mobihty window"; and fluid composition, flow volume and pathways. Thus, the type of sedimentary basin in which a sand was deposited strongly controls the cementation process. Sandstones of rift basins (arkoses) and collision-margin basins (litharenites) generally have only a few percent quartz cement; quartzarenites and other quartzose sandstones of intracratonic, foreland and passive-margin basins have the most quartz cement. Clay and other mineral coatings on detrital quartz grains and entrapment of hydrocarbons in pores retard or prevent cementation by quartz, whereas extremely permeable sands that serve as major fluid conduits tend to sequester the greatest amounts of quartz cement. In rapidly subsiding basins, like the Gulf Coast and North Sea basins, most quartz cement is precipitated by cooling, ascending formation water at burial depths of several kilometers where temperatures range from 60 o to 100 o C. Cementation proceeds over millions of years, often under changing fluid compositions and temperatures. Sandstones with more than 10% imported quartz cement pose special problems of fluid flux and silica transport. If silica is transported entirely as H4SiO4, convective recycling of formation water seems to be essential to explain the volume of cement present in most sandstones. Precipitation from single-cycle, upward-migrating formation water is adequate to provide the volume of cement only if significant volumes of silica are transported in unidentified complexes. Modeling suggests that quartz cementation of sandstones in intracratonic basins is effected by advecting meteoric water, although independent petrographic, isotopic or fluid inclusion data are lacking. Silica for quartz cement comes from both shale and sandstone beds within the depositional basin, including possibly deeply buried rocks undergoing low-grade metamorphism, but the relative importance of potential sources remains controversial and likely differs for different formations. The most likely important silica sources within unmetamorphosed shales include clay transformation (chiefly illitization of smectite), dissolution/pressure solution of detrital grains, and dissolution of opal skeletal grains; the most likely important sources of silica within unmetamorphosed sandstones include pressure solution of detrital quartz grains at grain contacts and at stylolites, feldspar alteration/dissolution, and perhaps carbonate replacement of silicate minerals and the margins of some quartz grains. Silica released by pressure solution in many sandstones post-dates the episode of cementation by quartz; thus, this silica must migrate and cement shallower sandstones in the basin or escape altogether. Some quartz-cemented sandstones are separated vertically from potential silica source beds by a kilometer or more, requiring silica transport over long distances. The similarity of diagenetic sequences in sandstones of different composition and ages apparently is the result of the normal temperature and time-dependent maturation of sediments, organic matter and pore fluids during burial in sedimentary basins. Silica that forms overgrowths is released by one or more diagenetic processes that apparently are controlled by temperature and time. Most cementation by quartz takes place when sandstone beds were in the silica mobility window specific to a particular sedimentary basin. Important secondary controls are introduced by compartmentalized domains produced by faults (e.g., North Sea) or overpressure boundaries (e.g., Gulf Coast Tertiary). Shallow meteoric water precipitates only small amounts of silica cement (generally less than 5% in most fluvial and eolian sandstones), except in certain soils and at water tables in high-flux sand aquifers. Soil silcretes are chiefly cemented by opal and microcrystalline quartz, whereas water-table silcretes have abundant normal syntaxial quartz overgrowths. Silica for silcrete cements and replacements comes from quartz, silicate minerals, and locally volcanic glass, in alluvium and bedrock.

0012-8252/89/$15.05

01989 Elsevier Science Publishers B.V.

70 INTRODUCTION Cementation is the most important process leading to the induration of sand to form sandstone. Sorby first documented the presence of quartz overgrowths in sandstones in 1880. Since then considerable interest and discussion has occurred concerning sources of silica for quartz cement and their relative importance, the time and depth of cementation of sands, the pathways and sources of water that transports the silica to the site of cementation, reasons for quartz precipitation, and other aspects of the cementation process that would be useful in understanding diagenetic processes in sandstones, including the ability to predict the distribution of quartz and other cements in the subsurface. Remarkably different interpretations have been made concerning the origin (Table I) and significance of quartz cement. For decades most of the reported data on quartz cement was qualitative, thereby permitting only general interpretations. More quantitative data are now being obtained, which sharply constrain interpretations. For example, because precipitation of quartz overgrowths is one of the first diagenetic events recognized in m a n y sandstones, m a n y workers report cementation to be an "early" event which took place before significant burial. Limited quantitative data indicate, however, that most quartz cement (except in silcretes) is introduced after sandstones have been buried 1 to 2 km and are subjected to temperatures greater than 50 C (Fig. 1; Table II). Depth of cementation can be calculated indirectly from fluid inclusion data (e.g., Haszeldine et al., 1984) and oxygen isotopic composition of quartz overgrowths (e.g., Land and Dutton, 1978), both of which yield values for the temperature of cementation. Using the present geothermal gradient or an age-corrected gradient, the depth of cementation can be calculated. In addition, pre-cement porosity values ( = intergranular volume or minus cement porosity) of a sandstone put constraints on the relative time of significant cementation by assessing the amount of porosity lost to compaction. Shallow, and generally early, cementation takes place when the intergranular volume is close to 40%, whereas deep, and

Earle F. McBride occupies the Wilton E. Scott Centennial Professorship in Geology at the University of Texas at Austin, where he started teaching in 1959. He was born in Moline, Illinois, U.S.A. in 1932. He received the A.B. degree from Augustana College, M.A. degree from the University of Missouri-Columbia, and Ph.D. degree from

the Johns Hopkins University. He has served as a visiting professor at the University of Kansas in the U.S.A. and University of Perugia in Italy. He is an Honorary Member of the SEPM, an Honorary Life Member of the Permian Basin Section of the SEPM, and a Fellow of the GSA. He has served the national SEPM as Councilor for Mineralogy, Secretary-Treasurer, and President. Professor McBride's research has been principally in sandstone petrology and diagenesis sedimentology of turbidites, the origin of bedded cherts, and the sedimentary history of Paleozoic rocks in the Marathon uplift of Texas. His address is: Department of Geological Sciences, University of Texas at Austin, P.O. Box 7909, Austin, TX 78713-7909, U.S.A.

71
0

TABLE I Sources proposed for silica in quartz cement in sandstones (1) Decomposition of feldspars (Sorby, 1880) (2) Silica precipitated from descending meteoric water; silica derived from silicate minerals in the zone of weathering (Van Hise, 1904) (3) From shales during compaction: silica derived from dissolution of quartz silt grains (Johnson, 1920) (4) From decomposed Equisetum (reeds) and Pteris (ferns) (Cayeux, 1929) (5) From siliceous sponge spicules in shales (Cayeux, 1929) (6) Precipitation at the sea floor from sea water (Krynine, 1941) (7) Pressure solution at grain contacts in sandstones (Waldschmidt, 1941) (8) Pressure solution at stylolites in sandstones (Heald, 1955) (9) Dissolution of opaline marine skeletal grains (diatoms, Radiolaria, sponge spicules; Siever, 1957) (10) Silica released from the hydration of volcanic glass formed opal cement (Swineford and Franks, 1959) or quartz cement (Fiichtbauer, 1974a) (11) Silica precipitated from descending meteoric water that is supersaturated with quartz before infiltration (Siever, 1959) (12) Silica released from silicates replaced by carbonates (Walker, 1960) (13) Silica released from the conversion of a smectite to illite in shales during burial diagenesis (Towe, 1962, using data from Burst, 1959; Siever, 1962) (14) Silica derived from the dissolution of eolian quartz abrasion dust (Biederman, 1962) (15) Silica released by naturally occurring silica complexes derived from organic acids (Evans, 1964) (16) Pressure solution of quartz grains in clayey siltstones (Fiichtbauer, 1967b) (17) Dissolution of detrital quartz grains in sand without pressure solution (Dapples, 1967) (18) Desorption of H4SiO4 from clays (Siever, 1971) (19) Pressure solution loss of detrital quartz silt in shale (Heald, 1955; Fiichtbauer, 1974b) (20) Silica derived from the dissolution of glacially produced abrasion dust (Whalley and Krinsley, 1974) (21) Abrasion solution of detrital quartz grains followed by precipitation of overgrowths (Amaral, 1975) (22) Detrital amorphous alumino-silicates that accumulate in muds and sands (Fisher, 1982) (23) Pressure solution of quartz and other silicates in deeply buried shales undergoing low-grade metamorphism (Land et al., 1987)

I-" 2

I000

8.~ !09 l
\4 ~
6+, 7 "9

zooo
12. W tm

,,\
,?

3000

4000

?" 15 I

14 - -- ~ J i~ 0

i00

TEMPERATURE (c)

Fig. 1. Depth-temperature plot of quartz cementation episodes. Solid lines show temperature-depth ranges for different formations where data are fairly well constrained; dashed lines are extrapolated from data points representing the minimum or maximum depth-temperature of cementation. Numerals in this figure identify the sources in Table I from which data were derived. Data for silcrete formations 1 and 2 are schematic. Not shown on the figure is data on the Norphlet Formation (Dixon et al., 1989), which plots between 5.4 and 6.8 km at temperatures between 130 o and 185 o C.

generally late, c e m e n t a t i o n o c c u r s a f t e r c o m p a c t i o n h a s p r o d u c e d a m u c h l o w e r intergranular volume. At present, however, no q u a n t i t a t i v e u n d e r s t a n d i n g of d e p t h versus c o m p a c t i o n exists. Q u a r t z c e m e n t is the m a j o r d e s t r o y e r of porosity and main control of reservoir quality in m a n y s a n d s t o n e s (e.g., T u s c a r o r a : Sibley a n d Blatt, 1974; T e n s l e e p : F o x et al., 1975; C o t t o n Valley: Bailey, 1983; T u s c a l o o s a : Smith, 1985; e x p e r i m e n t , P i t t m a n a n d Larese, 1987), b u t small a m o u n t s o f q u a r t z c e m e n t c a n r e t a r d or p r e v e n t significant p o r o s i t y loss by compaction. Hartshorne sandstones with abundant quartz cement have better reservoir q u a l i t y t h a n m a n y lesser c e m e n t e d s a m p l e s , the latter of w h i c h u n d e r w e n t m a j o r loss of porosity by intergranular pressure solution ( H o u s e k n e c h t , 1984), a n d C i s c o s a n d s t o n e s with 5% q u a r t z c e m e n t s u f f e r e d little loss of p o r o s i t y b y c o m p a c t i o n o f ductile g r a i n s in

i'O

TABLE II

Data on quartz cement: main depth of precipitation, duration of cementation, and temperature of cementation Duration (Ma) Age T ( C) Formation and locality Miocene Oligocene Reference Key to source in Fig. 1 Riezebos (1974) Thiry et al. (1988)

Depth (m)

Vol.% quartz cement range (avg) [ - 5] 0.2 [ < 40] 40 [ - 6] 1.6 3 < 40 75-80 40-50 68-94 F 55-75 - 80

100 < 100

500 500-1500 500-1500

locally tight 0 - 4 (0.6)

0-15

Netherlands Fontainebleau, Paris Basin Riley, Texas various Wilcox, Texas

Cambrian Paleozoic Eocene Cretaceous E. Jurassic Oligocene Pennsylvanian

9
8

900-1500

0-30 (13)

> 1000

17-25

Travis Peak, Texas "Jurassic"

10 11 3

> 1300 < 1700

0-12 0-19 (4)

McBride (1988) Fiichtbauer (1967) Fisher and Land (1986) Dutton and Land (1988) Haszeldine et al. (1984) Milliken et al. (1981) Mack (1984) Frio, Texas Cisco Group, Texas North Sea

1500-2000 > 50 - 60

1500-2000

0-31 (11)

Strawn, Texas,

Pennsylvanian

Lonoy and Akselsen (1985) Land and Dutton (1978)

1670-2100 [ < 24] > 6 90-100 Jurassic (80) Miocene

0 - 1 4 (5)

< 24

40-50

Miocene

McBride et al. (1988)

4 12 13

> 1800

"Miocene" offshore, La. G u l f of Thailand Travena and Clark (1986) McBride et al. (1987)

2000-2500

0 - 2 (1) * 0 - 2 0 (7) *

> 2440 > 2500


[ < 5]

0 - 1 8 (3)

3000

< 1

100 n 60 ( - 60) 100-120 (100) D 130-187 85-95 F 70-105 F 92-152 F Plio-Pleistocene Miocene Jurassic

L. Cretaceous Oligocene Pliocene

16 6 7 14 15

> 4000 < 4000 5400-6800

< 1 0 - 6 (0.6)

< 5 [ < 24]

Loucks (1978) Land (1984) Wilson and McBride (1988) Milliken (1985) G o l d (1984) Dixon et al. (1989)

Norphlet, G u l f Coast Pearsall, Texas Frio, Texas Pico-Repetto, California G u l f Coast Louisiana Norphlet, G u l f Coast N o r t h Sea N o r t h Sea Blakely Crystal Mtn. Ordovician

Matter et al. (1985) Matter et al. (1985) Nelson in Roedder (1979)

Note. Examples are listed in order of increasing depth of cementation; depth values are from original sources. All temperatures are based on oxygen-isotopic data, except where D = depth of occurrence ( T calculated by me from present geothermal gradient and depth of cementation) and F = fluid-inclusion data. I calculated temperature values in parentheses; these values were not reported in the original sources. Duration of cementation was derived either from burial curves that show the time span that the formations spent in the quartz cementation window or from the maximum ages of the formations. I calculated duration values in brackets.* = different cores

--,..I

74

contrast to less-cemented sandstones that lost nearly all porosity by compaction (Mack, 1984). When lithification processes are understood, the information can be used to predict where to find the best aquifers and hydrocarbon reservoirs. At least 23 possible sources of silica for cement have been proposed (Table I). The stable isotopes of Si apparently do not reflect any appreciable temperature fractionation and, along with trace elements, do not present any apparent means of chemically tracing Si from its source rocks. Therefore no definitive conclusions have been reached on the relative importance of the 23 hypotheses. Nevertheless, data obtained in the last decade yield some important constraints on the source of silica and the cementation process. This article reviews these and other aspects of the origin and distribution of quartz cement in sandstones. Cement is used here to indicate a mineral passively precipitated within the pores of a sand.
CHARACTER AND PRECIPITATION OF QUARTZ CEMENT

Quartz ooergrowths
Quartz overgrowths develop by the precipitation of silica directly from aqueous solution as well-ordered, low (alpha) quartz. The most c o m m o n form of quartz cement is an overgrowth, a syntaxial rim with the same crystallographic orientation and optical continuity as that of the detrital grain (Figs. 2-5). Overgrowths are one variety of Krynine's (1946) sedimentary (low-temperature) quartz and are megaquartz in Folk's (1968) size classification of sedimentary quartz. The ubiquitous occurrence of quartz overgrowths to the exclusion of new crystals (but see Stackelberg, 1968, p. 33; Fig. 5) indicates that the activation energy necessary to grow new crystals (i.e., homogeneous nucleation) is extremely large, and much greater than the activation energy needed to form overgrowths on detrital grains, the latter of which act as seeds (heteroge-

neous nucleation, cf. Putnis and McConnell, 1980, pp. 98-100). The manner of quartz overgrowth development is known from a large number of thinsection studies of quartz-cemented sandstones and specifically from the SEM studies of Waugh (1970) and Pittman (1972), and from experimental studies of Ernst and Blatt (1964), Heald and Renton (1966), and Paraguassu (1972). Overgrowths start as numerous tiny crystals that coalesce into a single large crystal with well-formed crystal faces if conditions of silica supply, time and space permit (Fig. 2). Pittman (1972) showed that overgrowths develop by either overlap and merging of individual similarly oriented subunits, or by envelopment of smaller subunits by one overgrowth that becomes dominant. Most incipient overgrowth subunits are fairly equidimensional, but some extremely elongate prismatic crystals can be developed (Waugh, 1970; Amaral, 1975; Fig. 4, this report). Long prismatic overgrowths are referred to as outgrowths. Overgrowths generally are in contact with the detrital seed grain only at widely spaced points, leaving a capillary pore between the cement and the majority of the detrital grain. Minerals, organic matter, or fluid trapped along the detrital grain margin forms the familiar "dust line" that can be seen in thin sections (Figs. 2, 3). In some sandstones, quartz precipitates later in the capillary pores and obscures the dust line (Pittman, 1972). However, nucleation of an overgrowth on a seed grain can not occur where the grain is coated with clay, iron oxides, organic matter or other minerals. As shown later, overgrowth development is prevented where foreign mineral coatings on quartz grains are thick (Fig. 4). Although quartz overgrowths form in optical continuity with detrital cores, overgrowths can differ from detrital grains in trace element composition and possibly structural state. Detrital grains generally luminesce red, blue or brown under excitation by electrons, whereas quartz overgrowths generally do not luminesce or luminesce faintly (Sippel, 1968;

75

Fig. 2. Quartz and chert grains with variable amounts of quartz cement. Quartz grains A and B have well-developed dust fines (black arrows) and overgrowthswith large pyramid and prism faces. Quartz grain C lacks visible dust fines and shows only incipient overgrowths.Chert grain D lacks overgrowthsand chert grain E has only tiny outgrowths (white arrows). Intergranular area is pore space. Miocene subsurface sample. Plain fight.

Zinkernagel, 1978; Matter and Ramseyer, 1985; Henry et al., 1986). Luminescence colors in quartz are attributed to various factors, including degree of lattice order, stress rate, the T i / F e ratio, A1 concentration, and occurrence of trace amounts of positively charged ions with small ionic radii (Sprunt, 1981; Matter and Ramseyer, 1985). Cathodoluminescent zoning of quartz overgrowths, attesting to some degree of heterogeneity in the overgrowths, has been reported by several workers (Blatt, 1979; A. Matter, oral commun., 1982; Land, 1984; Fisher and Land, 1986; Henry et al., 1986; and Suchecki and Bloch, 1988). Henry et al. (1986) employed cathodoluminescence and backscatter electron imaging on the electron microprobe to document the presence of complex oscillatory zoning or disrupted zoning in some overgrowths. Some of the zoning was the result of variable amounts of A1, commonly more abundant than 600 ppmw (which contrasts with most igneous quartz which contains only 100 to 200 p p m w A1; Nassau and Prescott, 1977; Henry et al., 1986). Elevated concentrations of A1 in quartz supresses cathodolu-

minescence (Ramseyer and Mullis cited in Matter and Ramseyer, 1985). This evidence of compositional zoning plus the presence of zoned fluid inclusions (Haszeldine et al., 1984), rare zoning visible in thin sections under the petrographic microscope (Austin, 1974), and zoning shown by etch patterns in H F (R.L. Folk, oral commun., 1987) indicate that cementation by quartz can procede episodically and from formation waters of variable composition. Many sandstones in the initial stages of cementation show that overgrowths form selectively at narrow constrictions, some of capillary size, between detrital grains. This phenomenon has been attributed to the decrease in activity of water (Tardy and Monnin, 1982) or to a gain of surface energy at detrital grain margins in tiny pores (Wollast, 1971). Both processes promote precipitation of silica. Overgrowths conform perfectly with the orientation of the crystal lattice of detrital grains, including those with undulose extinction or polycrystalline structure. Overgrowths conform with the orientation of each micro-

76

Fig. 3. Photomicrograph of quartz cement. A) Quartzarenite completely cemented by quartz. Dust lines are well-developed on nearly all grain margins. Plain light. Exotic pebble from olistostrome in Haymond Formation (Pennsylvanian), Texas. B) Crossed polars of above sample. Cement-cement contacts range from linear (hollow arrow) to curved. Triple junctions (solid arrows) are common in completely cemented sandstones.

crystal of quartz in chert grains also, although cementation proceeds slower or starts later on chert grains than on monocrystalline quartz grains (Heald and Renton, 1966; Fig. 2). If space permits, overgrowths develop smooth hexagonal and rhombohedral prism faces (Fig. 2), although thin isopachous coats of quartz occur in some silcretes cemented in the vadose zone (R.L. Folk, pers. commun., 1988). In the

competition for space in pores that become completely cemented, mutual or compromise boundaries develop that are generally irregular, although straight or concavo-convex contacts are locally c o m m o n (Figs. 3, 4). Straight overgrowth faces commonly form 120 angles and develop triple junctions with adjacent grains (Fig. 4). The relative growth rate of quartz varies with crystallographic direction

77

Fig. 4. Well-developed quartz overgrowths on quartz grains A and B and one large and several small outgrowths (arrow) on grain C. Uneven distribution of quartz cement reflects local, thick illite coats on surfaces of detrital quartz grains. Plain light. Norphlet Formation (Jurassic), Rankin County, Mississippi. and is fastest along the c-axis (Pittman, 1972). The internal structure of detrital quartz grains influences the volume of cement sequestered by individual grains. In the Jurassic Nugget Formation, for example, monocrystalline detrital quartz grains are three times more likely to possess well-developed quartz overgrowths than polycrystalline grains, and overgrowths in medium-grained sandstones show a preference, in decreasing order, for non-undulose, undulose, and polycrystalline grains (James et al., 1986). Several aberrant types of quartz cement exist, most being volumetrically small. The

Fig. 5. Example of exceptionally large, late-stage euhedral authigenic quartz crystal that is not in optical continuity with adjacent grains, at least in the plane of the thin section. These late-stage crystals typically form in secondary pores. Plain light. Norphlet Formation (Jurassic), Rankin County, Mississippi.

78 commonest of these aberrant forms in my experience is small prismatic quartz crystals that grow as outgrowths into large secondary pores as a late burial diagenetic event (Fig. 5). In thin section these crystals do not seem to be overgrowths. Another aberrant form of quartz cement occurs as a coating of micronsize quartz crystals either along or intergrown with authigenic clays. The coats range from < 2 ~m (Pittman and King, 1986) to 20 # m thick (Stackelberg, 1986). Cretaceous sandstones of offshore Gabon have a microquartz druse coating that is too thin ( < 2 ~m) to be seen in thin section (Pittman and King, 1986). Druse coatings on detrital quartz grains in the St. Peter Sandstone (Ordovician) of the midcontinental U.S. have been described as having botryoidal shape (Odom et al., 1979), as possessing well-formed crystals intergrown with illite (Stackelberg, 1986), and pictured as forming scale-like plates (Grutzeck, 1986). Yet another aberrant form of quartz cement is crusts 1 / 2 /~m thick by 2 ~ m wide of tiny quartz growths that lack crystal faces. These "turtle skin" crusts occur in m o d e m desert dune sands and are inferred to be the product of aging of opal crusts (Folk, 1978). Other thin ( < 1 /~m) quartz cement crusts that lack crystal faces have been reported to form on pits, cracks and edges of glacially fractured quartz grains (Whalley and Krinsley, 1974; Fillon et al., 1978). Precipitation is reported to start on some cleavage surfaces and later extend to stressed cracks and hollows, and precipitation is greater on smaller grains than larger grains, presumably because smaller grains have more disrupted crystal lattices (Whalley and Krinsley, 1974).
Other silica cements

ucts of surface silica diagenesis (Summerfield, 1986). Silcretes are summarized in a later part of this review. Several reported occurrences of chert-cemented sandstones exist, however, which were not identified as being silcretes (e.g., Friedman, 1954; Heald, 1956; Stalder, 1975; Amaral, 1975; Stanley and Benson, 1979; Hoholick, 1980; Sears, 1984), although only Heald and Sears studied marine sandstones. Sears (1984) described sandstones equivalent in age to the Miocene Monterey Formation of California which have extensive amounts of opal-CT (as lepispheres) and microquartz cement mixed with authigenic smectite. The authigenic silica phases were derived from opaline skeletal material in the sandstones and adjacent mudstones. Millot (1960, 1970) argued that as the concentration of "impurity elements" in groundwater decreases, the silica polymorphs listed above will precipitate sequentially. However, the scarcity of opal and other silica polymorphs as cements formed in the subsurface waters of diverse salinity indicate that there are more controls on polymorph precipitation than concentration of dissolved species. Various silica polymorphs are also present in sandstones in several Precambrian iron formations (e.g., Simonson, 1987), but these rocks apparently formed in sea water somewhat different from Phanerozoic sea water. DISTRIBUTION OF QUARTZ CEMENT IN TIME AND SPACE
General c o m m e n t s

Other polymorphs of silica that occur as cements in sandstones are opal-A, opal-CT, fibrous microcrystalline quartz (chalcedony, lutecite and possibly quartzine), and nonfibrous granular microcrystalline quartz ("chert"). Almost all occurrences of these polymorphs are in silcretes, indurated prod-

Quartz is perceived by some authors to be the most abundant cement and carbonate the second most abundant cement in sandstones (Tallman, 1949; Fiichtbauer, 1974a; Pettijohn, 1975, p. 447). However, this conclusion is based on the frequency of occurrence of cements in different formations, not on volume percent of the rocks. Quartz cement occurs in sandstones ranging in age from Precambrian to Pliocene/Pleistocene, and has even been reported in Holocene sands (Dap-

79 T A B L E III A m o u n t of quartz, c a r b o n a t e a n d clay cement in quartzarenite, arkose, a n d litharenite F o r m a t i o n (Age) location No. of samples Volume of cement (%) qtz carb. 2 2 -

Source

clay 3 6 1 1.0

Quartzarenites
Travis Peak (K), Texas, U S A M u d d y (K), Colorado, U S A F o u n t a i n (Perm), Colorado, U S A Sewanee (Penn), Georgia, U S A Tuscarora (Sil), Appalachians, U S A Kinnikinic (Ord) Idaho, U S A St. Peter (Ord) Illinois, U S A H a r d y s t o n (Cam), Pa., N.J., U S A H a r p e r s ( C a m - Precam?) Antietam (Cam-Precam?) Average: Y o w l u m n e (Mio), Calif., U S A Pyabwe (Mio) O k h m i n t a u n g (Olig) C h a t s w o r t h (K), Calif., U S A East Berline (Jur), Conn., U S A Chugwater (Tri), Wyoming, U S A Various (Tri) Conn., U S A F o u n t a i n (Perm), Colorado, U S A M i n t e r n (Penn) Colorado, U S A H a r d y s t o n (Cam), Pa., N.J., U S A Average: 255 32 5 32 185 12 88 2 20 46 14 14 Tr 21 30 17 16 3 1 11.6 Tr 1 Tr Tr Tr 16 1.9 D u t t o n (1987) Berg a n d Davies (1968) H u b e r t (1950) C h e n a n d Goodell (1964) Sibley a n d Blatt (1976) James a n d Oaks (1977) Stackelberg (1987) A a r o n (1969) Schwab (1971) Schwab (1970)

13 Tr 1
1.8

33 5 5 3 23 150 45 16 68 2

4 32 32 3 14 14 6.9 3.2 8 13

11 7 Tr Tr 2.0

Tieh et al. (1986) Satin (1963) Satin (1963) Link et al. (1984) H u b e r t a n d Reed (1978) Picard (1966) Krynine (1950) H u b e r t (1950) Boggs (1966) A a r o n (1969)

Litharenites
M a c i g n o ( O l i g - M i o ) , Italy Frio (Olig), Texas, U S A Molasse (Tert), G e r m a n y T o p o g o r u k (Tert), Alaska, U S A D i f u n t a G r o u p (K) Mexico B r o u g h t o n (Perm) Parry G r o u p (Dev), N.S.W. Australia Denbigh Grits/ E l a n d o n e r y G r i t s (Ord) Wales U.K. M a r t i n s b u r g (Ord), Appalachians, U S A Average: 14 22
-

1 1 Tr 0.2

1 19 20 1 10 8 5 7 7.9

9 6 11 2.9

7 81 18 19 5 21

Deneke and G u n t h e r (1981) N a n z (1954) F i i c h t b a u e r (1964) K r y n i n e in Payne et al. (1952) McBride et al. (1975) R a a m (1968) Crook (1960) O k a d a (1967) McBride (1962)

Note. Quartzarenite samples average > 95% detrital quartz; arkose a n d litharenite samples average more t h a n 25%
feldspar a n d rock fragments, respectively. Average values for each clan are unweighted for n u m b e r of samples per formation: qtz = quartz; carb. = carbonate.

ples, 1967, p. 325; Baltzer and Le Ribault, 1971; Folk, 1978). Based on a literature survey 39 years ago (Tallman, 1949), Early Paleozoic and Precambrian sandstones have four times as many quartz-cement-dominant formations as carbonate-cement-dominant formations.

The ratio is 2 : 1 for Late Paleozoic formations and 1 : 1 for Mesozoic and Cenozoic formations. Thus, older formations are more likely to have quartz cement. The Leder and Park (1986) model attributes this to a greater amount of time for fluids to have introduced

8O TABLE IV Factors influencing presence of amount of quartz cement (Many factors are related, some example references are given) Factor Grain size Influence Surface area Reference Renton and Heald (1966) Fiichtbauer (1967b) Stephan (1970) McBride (1984) Goldstein (1950) Fisher (1982) Land (1984) Haszeldine et al. (1984) (above citations) Renton and Heaid (1966) Wescott (1983) Mack (1984) Leder and Park (1986) Fiichtbauer (1974b) Fox et al. (1975) Krynine (1941) Siever (1959) Kantorowicz (1985) Hurst and Irwin (1986) Dutton (1986) Mack (1984) Wood and Hewitt (1984) Haszeldine et al. (1984) Cassan et al. (1981) Leder and Park (1986) Tallman (1949) Pettijohn (1957) Leder and Park (1986) Levandowski et al. (1973)

Grain size Permeability

Permeability and fluid flux Fluid flux

Fluid flux

Sequestering opportunity

Depth Depositional environment

Not cited Temperature Pore water composition (marine) Pore water composition (fluvial) Availability of biogenic silica Isolated versus connected sandstones

Focussed fluid controlled by rock geometry Convection

Nuidflux Huidflux

Time

Fluidflux

Structural position

Fluidflowpattem

silica. However, the distribution in part reflects the secular distribution of different types of sandstones. Quartzarenites and other quartz-rich sandstones occur chiefly in cratonic sequences, which also tend to have the greatest amounts of quartz cement (Pettijohn, 1957; Pettijohn et al., 1987, p. 447; Table III). Sandstones of tectonically more active sedimentary basins are less quartz rich, tend to have less quartz cement but have greater amounts of clay minerals, carbonates

and other cements (Table III). In mineralogically immature sands, pore waters are more likely to contain higher activities of a wide variety of dissolved components such that dissolved silica may form clay minerals, zeolites, and other silicate species (L.S. Land, oral commun., 1985; Pettijohn et al., 1987, p. 447). The greater abundance of quartz cement in quartz-rich sandstones has also been attributed to the greater opportunity for deriving silica from detrital quartz grains from such

81 TABLE V Textural selectivity of quartz cement Formation Cardium Mountain Park Keuper Frontier Hosston M. Buntsandstein M. Buntsandstein Ivishak Norphlet Tensleep Hartshorne Age L. Cretaceous L. Cretaceous Triassic L. Cretaceous E. Cretaceous E. Triassic E. Triassic Permo-Trias. Miocene Jurassic Pennsylvanian Pennsylvanian Quartz selective to Finer beds Finerbeds Finer beds Finer beds Finer beds Finer beds Finer beds Finer beds Finer beds Coarser beds Coarser beds Coarser beds Coarser beds Reference Thomas and Ohver (1979) Mellon (1964) Fondeur (1964) Heling (1965) Tillman and Almon (1979) Fielder et al. (1985) Fiichtbauer (1967b) Stephan (1970) Melvin and Knight (1984) Travena and Clark (1986) McBride (1984) Fox et al. (1975) Houseknecht (1984)

sands (Blatt, 1979, p. 147), although, as discussed later, this remains controversial. The amount of quartz cement is highly variable in different formations, in different beds within the same formation, and in different parts of the same sample. Very few sandstones contain quartz as the only cement, and few sandstones show all pores completely filled by quartz cement. In fact, sandstone beds that are entirely cemented by quartz are a rarity. F e w diagenetic settings produce sufficient silica to completely cement a given sandstone, which is fortunate for the petroleum geologist! In fact, tightly quartz-cemented sandstones present a problem in explaining the source of the huge volumes of imported silica. Some of the reasons for variation in quartz cement abundance (Table IV are discussed later in this review.

Textural selectivity of quartz overgrowths


Quartz overgrowths commonly show selectivity for beds or laminae of a particular grain size within a formation. Case studies in the literature suggest that twice as many formations show a selectivity of quartz cement for finer-grained beds as opposed to selectivity for coarser-grained beds (Table V). However, in m y experience I have found the preference

of quartz cement for coarse-grained beds to be so pervasive that I listed it as one of the "rules" of sandstone diagenesis (McBride, 1984). In the few formations where I found quartz cement to be predominant in the finer-grained beds, it is because calcite cement had preceeded quartz cement and has preferentially cemented the coarser beds. Thus, quartz cement is dominant in some finer-grained beds b y default, but this does not apply to all formations. The experiments of Heald and R e n t o n (1966) indicate that well sorted, coarser sands became cemented faster than finer sands with freely circulating cementing fluids, but finergrained beds cemented faster than coarser beds when the flux was the same for both beds. These authors and Fiichtbauer (1967a) and Stephan (1970) attributed the greater abundance of nucleation sites (surface area) in finer-grained sands as the reason for their selectivity of quartz cement. However, several studies indicate that fluid flux is one of the most important factors influencing the volume of quartz cement in sandstones. Wescott (1983) reported quartz to have a preference for clean, well-sorted, "high-energy" sandstones; Goldstein (1948, p. 120), Fisher (1982), and Land (1984) found it to be selective to more permeable beds; and Mack (1984) found

82 it to be selective to beds that transmitted the largest volumes of compaction-derived water. The greater permeability of coarser beds dictates that they can provide the largest fluid flux and, thus, a possibility of sequestering more silica for overgrowths than finer beds. Quartz grain structure (undulosity, polycrystallinity; James et al., 1986) can influence the sequestering of greater amounts of quartz cement in finer-grained beds because they have fewer polycrystalline and undulose grains than coarser-grained beds (Anderson and Picard, 1971). The degree of influence by this process must be small in most formations, however. Too few published case studies have constrained controlling variables sufficiently to provide an adequate data base on which to make predictive rules at present. Some formations have a grain-size selectivity for the amount of quartz cement present, but grainsize selectivity is not consistent among different formations, and several processes are known that can explain observed selectivity relations. and possibly 6 km. The data plotted in Fig. 1 are strongly biased toward Cenozoic formations (75% of total) and particularly biased toward the Gulf Coast basin (56%). The Gulf Coast basin is an actively subsiding, deep, passive-margin basin with local geothermal gradients greater than 60 C / k m and with a predominance of mineralogically immature sandstones that have an average of only 5% quartz cement. Pattern, rates, and amounts of quartz cement in sandstones from other types of basins are likely to be different from the Gulf Coast basin. The relation between burial depth and amount of quartz cement in a formation or basin is poorly known because many studies are restricted to outcrop samples, and most studies of subsurface samples have been restricted to a narrow depth range biased toward the oil window. The few studies with sufficient control to examine depth-cement trends yield diverse results. Amounts of quartz cement in Pennsylvanian sandstones in the Illinois basin and adjacent areas of the midcontinent U.S. have no relation with depth or with geographic location, structural position, or amount of pressure solution (Siever, 1957). The St. Peter Sandstone (Ordovician) in the northern Illinois basin displays no relation between quartz cement and depth (but see below), although only deeper sandstones are completely cemented by quartz (Hoholick, 1980; Hoholick et al., 1984). The Lyons (Permian) Sandstone in the Denver basin shows no correlation of quartz cement with depth over the depth range of 5800-9200 ft. (Levandowski et al., 1973), and Cretaceous sandstones of offshore Brazil show no depth trend with quartz cement over the depth range from 1.5 to 4.5 km (Chang, 1983). However, several workers report a general increase in amount of quartz cement with increasing depth in other formations. Fiichtbauer (1974a) found a linear increase in the number of quartz grains with cement over a depth range of 2500 m in Mesozoic sandstones in W. Germany (Fig. 6). The Tensleep and Weber (Mississippian) sandstones of Wyoming are

Depth trends
Two depth trends are examined here. One is the depth range over which cementation is most important, the other is the variation of amount of cement with depth. Both help to constrain quartz cementation history. A cross-plot of the depth of major cementation events versus temperature of 2 silcrete and 14 subsurface formations is shown in Fig. 1. The depth/temperature ranges are not well constrained for many of the formations plotted. In addition, the temperatures of cementation inferred to exist at a given depth (e.g., 2 km) span a range (45 o C) greater than expected for normal geothermal gradients. How much of this variation is real rather than the result of erroneous data is uncertain. Nevertheless, the plot indicates that little cementation occurs shallower than 1 kin, that the most active zone of cementation is between 1 and 2 km (over temperatures of 40 -90 o C), and that cementation can occur as deep as 4

83

500
m

'...
O

-35
-30
O

LIO T IiI 123

1000-

1500~ 2000X

-25 ~)

.~ 2500-

-20

,b

2b

36

46

sb

60

P R E T G OF QUARTZ GRAINS E CNA E WITH O E G O T S VR R WH

Fig. 6. Percentage of quartz grains with overgrowths versus maximum depth of burial for Dogger Beta sandstones (Jurassic), western Germany. Each dot represents the average of numerous samples for which the estimates of the number of quartz grains possessing three different degrees of overgrowths were made on loose grains in transmitted light. (After Fiichtbauer, 1974a, fig. 3-50.)

deeply buried in local basins but are exposed on the flanks of basement highs. The sandstones have less than 10% and locally less than 5 % porosity in the basins, but more than 15% porosity on the high parts of uplifts (Fox et al., 1975). Porosity in these formations is largely controlled by the amount of quartz cement. Higher temperatures are interpreted to result in greater amounts of quartz cement, but the source of silica was not identified. Quartz cement is reported to increase with depth (values not reported) along with chlorite and illite, in Eocene sandstones in Venezuela (Ghosh and Aguado, 1985), and in Vicksburg sandstones (Oligocene) of the Texas Gulf Coast between depths of 2000 and 5000 m (Fig. 8c). In the southern Illinois basin, the St. Peter Sandstone shows a weak but positive correlation of quartz cement with depth (Stackelberg, 1986). In the Gulf Coast basin, certain cementdepth trends occur in various formations. The most common pattern of cementation is one of trivial or small amounts of quartz cement until a threshold depth is reached, and then a relatively large increase in cement occurs over a depth range of approximately 1 / 2 km. Lit-

tie or no increase occurs at depths greater than the threshold depth (Figs. 7, 8). (Apparent small increases in quartz cement volume with depth [e.g., Miocene, Vicksburg] may be an artifact of compaction. During compaction, the volume of a rock decreases, resulting in a relative increase in all mineral phases, including cement.) The volume loss in Miocene and younger rocks is the result of loss of intergranular volume entirely, because pressure solution is not important. The coincidence of the quartz cementation threshold depth with the top of overpressured zone in the Frio Formation and onshore Miocene sandstones suggest a causal relationship (Land, 1984; Land et al., 1987). The suggestion was made, for the Frio especially, that cementation took place just below the top of the overpressure zone. Deeper sandstones with quartz cement are interpreted to be those that underwent subsidence after cementation. Wilcox s a n d s t o n e s c o n t a i n significant amounts of quartz cement above the top of overpressure zone, but this may reflect the lowering of the overpressure zone from pressure dissipation subsequent to cementation. Convecting water, perhaps derived from deep

I00

200

Travis Peak 14

I Cotton ? Valley I

rr.;
Plio/Pleist0cene ~ I Miocene 07

No,oh,et'4 '17
7

(3W 123 4

I00

200

AGE (ma)

Fig. 7. Depth of cementation by quartz versus depositional age of some Gulf Coast Tertiary and Mesozoic sandstones. Numerals are the average amount of quartz cement in the respective formations. For the last 100 Ma the depth of cementation apparently increases. Sources given in Table II.

84
8

(A)
n=44 -3

2;,,
4

I0.

(B)

: 576

12.

0 _o

4
14
II
e e ~ e=,

6
elill

_
5~

oS 8

-'r I-.(3. hi t:3

16~

o
=

..:..

3.

ql

x
I0

.:%

18-

-r,T 'i.-.....;...

. :*:: " '


""

20-

.,. ,2-

11~'4-'~"
.11#1.",, ipe l i e I

22
0

I
I

I
2

I
3

I
4

I
5

QUARTZ

CEMENT

(%)

16-

-.--...., :;y....::.:.

"

e e

ee

i
(C)
n = 203 8o 18

eoe

QUARTZ

CEMENT

(%)

I " "
4

(D)
, e l l I I I

~ IC 0
0 I

6i
II
I

I :""1
l.ql.l*

;,. '~ " ..

n = 207 2

e l

::"

l : l ' : "
I 13(:3
e

0
-r

~,.'wIb

te

e ..

f w,...-.:

.'"

3~
Q. i,i 121

0 e e

4
O 0 0 Q

~124

0 0

o o o 0

14 @

,
QUARTZ
CEMENT (%)

16 I0

i 2

l 4 QUARTZ

J 6

1 I 8 I0 12 CEMENT (%)

Fig. 8. Quartz cement versus present burial depth for four Gulf Coast Tertiary sandstones. A) Miocene, onshore Louisiana, (Gold, 1984). B) Free (Oligocene), Texas, (Land, 1984). C) Vicksburg (Oligocene), South Texas, (J. Wilson, unpubl.). D) Wilcox (Eocene), central Texas, (Stanton, 1977, and Fisher, 1982).

within the Gulf basin, that reached the top of the overpressure zone is a convenient supply of silica to call upon (Land, 1984; Land et al., 1987). This hypothesis needs further evaluation, because the age of overpressuring is unknown and large-scale convection in this part of the Gulf basin remains to be proven (Blanchard and Sharp, 1985).

Mesozoic sandstones in the Gulf Coast possess from 5 to 15% more quartz cement than Tertiary sandstones (Table VI). They are generally more quartz-rich than Tertiary sandstones, have a different stratigraphic setting (they are interbedded with carbonates and evaporites and are closer to basement), and possibly were invaded by hotter waters

85 TABLE VI Data on quartz cement in Gulf Coast sandstones Age/Unit Q :F :R Major cements (Cal, Qtz) Cal, Qtz, Kaol Cal, Qtz (Chl, Kaol) Cal, Qtz, Kaol Cal, Chl, (Qtz) Ank, Qtz, KaoI Qtz, Ank, Cal, Chl Qtz, (Ill, Chl, Cal) Qt__zz,Dol, Feld, Chl Qtz, Hal Anhy, Carb 14 14 7 Quartz cement (%) < 1 0.7 5 2.3 Depth (m) of impt. qtz. cmt. > 4000 > 4500 1670 to 2100 2500 to 3000 2700 1800 to 2500 2400 to 3500 900 to 1500 < 2500 2300 8180 qtz. cmt. N.A. N.D. + 34 + 31 Temp. of precip. ( o C) 9 ? 40-50 60 Reference

Plio-Pleistocene Miocene (Onshore) Miocene (Offshore) Oligocene: Frio Fm. Oligocene: Vicksburg Fm. Eocene: Wilcox Fm. Cretaceous: Tuscaloosa Fm. Cretaceous: Travis Peak Fm. Jurassic: Cotton Valley Jurassic: Norphlet (Miss.)

65 : 18 : 17 83 : 9 : 8 (variable) 86 : 10 : 4 25 : 30 : 45 to 65 : 20 : 15 18 : 39 : 43 65 : 15 : 20

Milliken (1985) Gold (1984) Land et al. (unpublished) Land (1984), Milliken et al. (1981) J. Wilson (1987), pers. commun. Stanton (1977), Fisher and Land (1986) Suchecki (1983) Dahl (1984)

2 1.6

? + 25

9 - 80

90 : 0 : 10 95 : 4 : 1 84 : 8 : 8 75 : 15 : 10

+ 14 to 17 + 19 ? + 18 90-100 55-75

Dutton (1987) Bailey (1983) McBride et al. (1987)

Note Cements: Cal = calcite, Qtz = quartz, Kaol = kaolinite, I11 = illite, Hal = Halite, Dol = dolomite, Ank = ankerite, Carb = unspecified carbonates, Chl = chlorite, Anhy = anhydrite. Cements underlined are dominant; cements in parentheses are minor. Other abbreviations: Impt. = important, Cmt. = cement, Feld = feldspar. After Sharp et al. (1988)
e a r l y d u r i n g their d i a g e n e t i c h i s t o r y ( M c B r i d e et al., 1987; L a n d et al., 1986; D u t t o n , 1987). T h e r e l a t i v e i m p o r t a n c e o f these d i f f e r e n t c o n d i t i o n s is u n c e r t a i n at this time. supported by textural evidence. Except for silcretes, T e r t i a r y s a n d s t o n e s p o s s e s s s m a l l amounts of quartz cement, and only four r e p o r t s o f H o l o c e n e q u a r t z c e m e n t exist, n o n e of which describe volumetrically important cement, and several of which are of q u e s t i o n a b l e validity. S a n d s t o n e s o f t h e G u l f C o a s t in g e n e r a l h a v e m o r e q u a r t z c e m e n t w i t h i n c r e a s i n g age ( T a b l e VI), b u t b u r i a l depth, temperature history, stratigraphic position, lateral c o n t i n u i t y a n d o t h e r f a c t o r s s e e m t o b e c o n t r i b u t i n g f a c t o r s a l s o ( L a n d et al., 1987; S h a r p et al., 1988). A much larger data base on sandstone framework and cement composition than was a n a l y s e d b y T a l l m a n (1949) m u s t b e ex-

Secular trend
T a l l m a n ' s (1949) d a t a i n d i c a t e t h a t o l d e r sandstones have more quartz cement than y o u n g e r ones. Is this a real s e c u l a r t r e n d , o r is the c o m p a r i s o n b i a s e d b y c o m p a r i n g older, cratonic quartzarenites with younger, immature sandstones (litharenites and arkoses) from rapidly subsiding basins? The argument t h a t o l d e r s a n d s t o n e s h a d m o r e t i m e f o r silica to r e p l a c e earlier c a r b o n a t e c e m e n t s is n o t

86 amined to evaluate the temporal significance of quartz cement. This is not easily accomplished, because few petrographic reports provide complete data on framework, cement and porosity.
Cementation at shale contacts

Enrichment of quartz cement at shale contacts has been found in some formations, providing evidence that some silica probably came from interbedded shales. Fiichtbauer (1967b) found a major quartz-cement enrichment zone 1.5 m thick at the tops and bases of sandstones interbedded with shales. Because of the distribution of quartz cement, only sandstone beds thicker than 3 m display decent reservoir quality. He concluded that cementation at both the top and bottom of beds precluded the introduction of silica from water expelled from shales during compaction and argued that the silica was introduced from shales by diffusion. Frio (Oligocene) sandstones of the Texas Gulf coast have been studied by several persons who report diverse findings. Moncure et al. (1984) found quartz cement enriched five-fold over average cement abundance at the lower contact of one sandstone bed examined in detail, Sullivan (1988) found only one-third of more than a dozen contacts to have quartz cement enrichment in sandstones, and Lindquist (1976) and Land (1984) found no systematic quartz cementation near shales. Additional work is needed to explain these inconsistent relations.
Diagenetic sequences and models

Paragenetic sequences of authigenic minerals and diagenetic events, such as dissolution and albitization, have been established for dozens of stratigraphic units of various ages in many different sedimentary basins chiefly over the past 15 years (e.g., Schmidt, 1976; Hurst and Irwin, 1982; Franks and Forester, 1984; Loucks et al., 1986). These sequences differ somewhat in specific details, but the sequential order of events, including precipi-

tation of quartz overgrowths as one of the first major events, is remarkably similar among the different formations (McBride, 1982). Surdam and Crossey (1987) developed a temperature-dependent model by considering the evolution of a system of aluminosilicate minerals, carbonate minerals, organic acids and CO 2 as the components passed through temperature windows of 80 to 120C and 120 to 200 C. They expanded the model to include a broader range of diagenetic reactions and temperature zones, and presented an idealized diagenetic sequence (Surdam and Crossey, 1987). In their model, quartz cementation is shown to occur in two episodes, one between 20 and 80 C (chiefly between 40 and 60 C) and one at temperatures greater than 160C (Surdam and Crossey, 1987, fig. 10). Data summarized in Fig. 1 of this report suggest that quartz cementation rarely occurs below 40 C, and that it takes place over a greater range of temperatures than suggested by the Surdam and Crossey model. The latestage episode of quartz cement of their model is at considerably higher temperatures than most of my data or experience support. In spite of these contrasts, the presence of a general diagenetic sequence is supported by considerable data. I suggest that much of the data summarized herein can be explained if many burial diagenetic reactions are considered to be the evolutionary product of both time and temperature, instead of temperature alone as proposed by Surdam and Crossey. This approach seems necessary to explain the wide range of cementation temperatures reported for quartz cement. Thus, the similarity of diagenetic sequences in sandstones of different composition and ages can be interpreted to be the result of the normal interaction of sediments, organic matter and pore fluids during burial in a sedimentary basin. The rocks, organic matter and aqueous pore fluids undergo compaction, heating and aging--all parts of the normal maturation process of basin fill. It seems likely that an optimum time-temperature stage exists during which

87 relatively large amounts of silica are released by various diagenetic processes. Quartz cementation may be expected to be most vigorous when sandstones, during their subsidence history, reside within this "silica mobility window" or when formation water charged with silica from the window migrates upward to cooler depths. Cementation by quartz, then, is not a chance event but one that is predictable. Leder and Park (1986) modeled quartz cementation for quartz-rich sandstones a n d found good agreement between predicted and actual amounts of cement for numerous ancient formations. They reported the important variables in decreasing importance are: (1) burial rate; (2) age; (3) initial porosity; (4) basin size (dip angle of beds); (5) fluid drive mechanism; (6) initial permeability; and (7) geothermal gradient. Their model requires upward flow of cementing waters at a rate of - 10 c m / y r with fluid drive by a combination of meteoric head and convection. The model ignores compaction and assumes cementation is a continuous, if uneven, process. Numerous studies document the importance of compaction, even in quartz-rich sandstones (e.g., Wilson and Sibley, 1978; McBride, 1987a; Scherer, 1987), and petrographic evidence summarized herein indicates that cementation by quartz generally is not a continuous process. These shortcomings suggest that the Leder and Park model should be used with caution. primary control on the characteristics of a sedimentary basin and the composition of detritus deposited in it. The average amount of quartz cement in 10 representative quartzarenites, arkoses and litharenites is 12, 2 and 0.2%, respectively (Table III). As mentioned previously, mineralogically immature sandstones (arkoses and litharenites) have unstable grains whose hydrolysis during diagenesis yields a greater activity of metal ions in formation water. The availability of these cations promotes the precipitation of authigenic clays, zeolites and other silicates as cement in arkoses and litharenites, whereas quartz cementation is more likely in quartzarenites. The relation between tectonics and source area is well established (Krynine, 1948; Pettijohn, 1957; Dickinson, 1985). Other controls on cementation are factors that affect the composition, flux, and flow paths of groundwater in a basin, the compaction history of the sediments, and the thermal history of rocks and fluids. These factors differ considerably in different types of basins, although most effects are known only qualitatively. For example, the rapid burial and compaction in a collision-margin basin of litharenites composed of abundant ductile grains will result in greatly reduced sandstone permeabilities, such that the hux of quartz-cementing fluids is retarded. In contrast, quartzose sandstones in an intracratonic basin will retain considerable permeability after burial, permitting long-term cementation by quartz. Scenarios for chemical diagenesis and hydrological evolution of different types of sedimentary basins have appeared recently. Siever (1979) and Bjorlykke et al. (1979) review the general diagenetic processes of major types of basins, whereas Dutton and Land (1985) and McBride et al. (1987) describe cementation by quartz and other minerals specific to a basin adjacent to a marginal uplift and in a rift basin, respectively, and Land et al. (1987) and Sharp et al. (1989) describe cementation in Gulf Coast passive-margin deposits. In ad-

Tectonic setting
Several lines of evidence show that there is an important relationship between the amount of quartz cement in a sandstone, its framework composition, and the type of sedimentary basin in which it was deposited. Framework composition, in part, influences whether quartz or other authigenic silicates precipitate in the pores of a sandstone; and the hydrologic and thermal history of a basin influence the time of cementation and the pathways of the cementing fluids. Tectonic setting is a

88

dition, hydrologic modeling of intracratonic and foreland basins (Cathles and Smith, 1983; Garven and Freeze, 1984a, b; Bethke, 1986) delimits the relative importance of meteoric and compactional water as cementing agents. It is apparent that the Gulf Coast basin, from which we have many case studies, is an inappropriate model for quartz cementation in foreland basins. The Gulf Coast basin hydrology is strongly influenced by compactional water, possibly undergoing convection, by a hydrologic boundary between normally pressured and overpressured regimes, and by abundant growth faults and local salt domes (Sharp et al., 1988). In contrast, the Alberta basin, a foreland basin, evolved from a compaction-driven groundwater flow system to a gravity-driven system dominated by cross-formational flow of meteoric water (T6th, 1978; Garven and Freeze, 1984a, b). More individual case studies from diverse types of basins are needed to sort out the role that the many variables play.
CONTROLS ON PRECIPITATION

General c o m m e n t s

Precipitation of quartz in sandstones can take place when the solubility product of quartz is exceeded and where detrital quartz grains are available as seeds. Kinetic factors are important in controlling precipitation, at least at low temperatures and low levels of quartz supersaturation. Although most surface and shallow subsurface meteoric water is supersaturated with respect to quartz (Livingstone, 1963) and is capable of being a cementing agent, there is no evidence that Holocene cementation by quartz is important. The local abundance of quartz cement in some silcretes suggests, however, that conditions can exist at or near the surface that permit large-scale precipitation of quartz (e.g., evaporation of ground water). Abnormally high levels of supersaturation, controlled by a m o r p h o u s phases, or the role or organic complexes in

transporting silica may also be controlling factors. The evidence, strongly biased toward passive-margin basins, is that quartz cement forms in the deeper subsurface ( > 1 or 2 km) when ascending hot formation water cools and silica reaches saturation. Silica lost by pressure solution in Upper Carboniferous sandstones in Germany can be accounted for as cement in overlying Lower Permian sandstones, supporting the upward migration theory (Fiichtbauer, 1974a, p. 139), and uncemented shadows in the Tuscarora Formation (Silurian) suggest cementation of quartz was by upward-directed fluid flow (Heald and Anderegg, 1960). The solubilities of quartz and other low-temperature silica polymorphs increase with temperature (Siever, 1959). A certain level of supersaturation of silica with respect to quartz may be required to trigger cementation in the subsurface also (e.g., Land, 1984, assumed a supersaturation value of 50 p p m to initiate cementation in the Frio Formation of the Gulf Coast). Change in p H has been suggested as a reason for precipitation also. Levandowski et al. (1973) suggested that high-pH alkaline brines saturated with silica precipitated quartz upon mixing with low-pH meteoric water in proximal facies of the Lyons (Permian) Sandstone in the Denver basin. However, the existence of subsurface brines with high pH derived from evaporites has not been documented (J. Warren, oral commun., 1987). Silica that must be transported as organic complexes would be sensitive to pH changes, but the abundance of such complexes and their chemical behavior are undocumented. Of the many potential sources of silica for quartz cement, most release silica at depths greater than the depth of cementation. The temperature range between 80 o and 120 o C is one of highly active chemical diagenesis (Surdam and Crossey, 1987), where silica likely is released by pressure solution, clay transformations in shale, and other mechanisms. Upward migration of silica-bearing water more than a kilometer seems likely for

89 most cementation episodes. Greater distances of fluid transport are required if metamorphism is a source of silica or if convection is operative (see following sections). Other factors promoting precipitation of quartz that have been cited include the lowering of pore pressure, mixing silica-saturated fluid with more saline formation waters at a given temperature or pressure, and osmosis through a clay-rich shale layer (Leder and Park, 1986). The relative merits of these mechanisms are difficult to test. The apparent selective cementation of Eocene Gulf Coast sandstones near the top of the overpressure zone (Land, 1984; Land et al., 1987) occurs at an interface where important differences in temperature, pressure and fluid composition may exist. Migration of water saturated with silica across the overpressure-normal pressure boundary should not result in significant precipitation of quartz cement if quartz behaves like amorphous silica. Experimental data (Willey, 1974) indicate that the difference in solubility of amorphous silica in water that passes from an overpressured zone (0.7 psi/ft.) to a normally pressured zone (0.465 psi/ft.) at 3048 m is only 0.4 ppm. How long does it take to introduce the volume of quartz cement in a particular sandstone, does cementation proceed from water of uniform composition, and is quartz cementation a separate event in diagenetic sequences? The answers to these questions are incomplete at this time. The absolute age of buried Tertiary formations places a maximum time limit on the cementation process, but relatively few of these formations have more than several percent quartz cement. Few workers have been able to constrain cementation events in older formations with abundant quartz cement. If the burial curve for a formation can be established and the temperature range of quartz cementation can be bracketed from oxygen isotopic composition of overgrowths (e.g., Land and Dutton, 1978; McBride et al., 1987) or from fluid inclusions in overgrowths (e.g., Haszeldine et al., 1984; Matter et al., 1985), then the timing of cementation can be calculated. Data from various studies are summarized in Table II. The length of time required to cement subsurface sandstones with at least 5% quartz cement ranges from as short as 1.6 to perhaps more than 50 Ma, with more than 6 Ma needed for most formations. If silcrete is ignored, sands less than 6 Ma have only minor amounts of quartz cement, and most Quaternary/Tertiary sandstones were deeply buried before undergoing partial cementation. Many papers on diagenetic sequences report quartz cementation as a separate event, which gives the impression that cementation proceeded under uniform chemical conditions. Whereas this scenario is possible, evidence is mounting that quartz cementation proceeds from formation water of variable composition and that cementation can overlap other diagenetic events. Evidence based on variable overgrowth morphology and on compositional zoning in overgrowths has been discussed previously. There are some reports of coprecipitation of quartz and other authigenic minerals, and there are some examples where cementation by quartz was interrupted by or overlapped an episode of calcite cementation (e.g., Land et al., 1987; McBride et al., 1988).

Inhibition of quartz cement


With few exceptions (e.g., veins, petrified wood), quartz that precipitates in pores requires detrital quartz seeds on which to grow. The average terrigenous sandstone has sufficient quartz grains to act as the necessary nuclei to promote cementation, but oolite and carbonate skeletal sands and glauconite sands may lack such nuclei and escape cementation by quartz (but they may be replaced by quartz). Sands with only a few quartz grains in the framework population may develop patchy quartz cement with odd overgrowth patterns. Several workers reported that mineral coatings on detrital quartz grains can inhibit

90 quartz overgrowths by isolating detrital grains from water capable of precipitating quartz overgrowths (e.g., Fiichtbauer, 1967b; Pittman and Lumsden, 1968; Pittman; 1972; and Heald and Larese, 1974). In the Gifhorn trough of West Germany, sandstone bodies with chlorite-coated quartz grains have the same average porosity as sandstones whose quartz grains lack chlorite coatings but are 300 m shallower (Fiichtbauer, 1967b, p. 354). In these sandstones, chlorite apparently preserved primary porosity to depths greater than would have occurred otherwise. The extraordinary porosity of the deeply buried Tuscaloosa (Cretaceous) sandstones of Lousiana is attributed by several workers to the quartz-inhibiting action of unusually thick chlorite coatings (Thomson, 1978b; Larese et al., 1984; Smith, 1985), although this interpretation is disputed by some (e.g., Franks, 1980). The effectiveness of different minerals in inhibiting quartz cement, in order of decreasing effectiveness, is chlorite > illite - chert = carbonate > hematite (Heald and Larese, 1974). Chlorite coatings include several minerals of the chlorite group, such as chamosite and sudoite. Grain coatings have various origins, including emplacement of clays by infiltering near-surface groundwater (Crone, 1974), formation of hematite as a near-surface oxidation product of Fe-bearing silicates (cf. Walker, 1967), clay cements introduced during burial (cf. Wilson and Pittman, 1977), and early diagenetic rim calcite cement (cf. James, 1985). The thickness and continuity of clay coats on detrital quartz grains also influences the effectiveness of the coats in inhibiting overgrowths. If clay coats are discontinuous, quartz cement may form skinny prismatic attachments (Fig. 4; termed outgrowths by Lowry, 1956). In some formations, grain coatings that were effective inhibitors of overgrowths early in the cementation period were removed before cementation ceased. Clues in thin section to the present (Fig. 4) or previous existence of gram coatings on quartz grains are summarized by Heald and Larese (1974). Consensus exists that entrapment of hydrocarbons in a sandstone essentially inhibits cementation by quartz, as well as other cements (e.g., Lowry, 1956; Fiichtbauer, 1967b; Stephan, 1970; Prozorovich, 1970; Yurkova, 1970; Thomas and Oliver, 1979). Some workers maintain, however, that cementation by quartz continues slowly after oil entrapment by diffusion of silica along the water film that coats all sand grains (e.g., Fiichtbauer, 1967b; Longman, 1976). It is uncertain, however, what volume of quartz might be precipitated this way. Oil drops in thin quartz overgrowths in oil-saturated sandstones of the Bromide Formation (Ordovician) of Oklahoma are reported to have been trapped during cementation by diffusion in the water film (Longman, 1976).

Environment of deposition
The environment of deposition influences cementation by quartz directly by controlling the composition of early pore water and indirectly by affecting the texture of sand (grain size, sorting), and, hence, its permeability, and the geometry of the sand, which affects its character as a fluid conduit (Stonecipher et al., 1984). These latter aspects have been discussed previously. Several workers claimed a direct relation between environment of deposition and quartz cement, but few arguments are well substantiated. Krynine's (1941) argument that sandstones are strongly cemented at the sea floor cannot be supported. Siever (1959) suggested that fluvial sands may undergo early cementation by quartz when river water and shallow ground water, both saturated with respect to quartz (Siever, 1962; Livingstone, 1963), invade fluvial aquifers. Fluvial facies of the Ravenscar Group (Jurassic) of the U.K. contain more quartz cement than other facies, which Kantorowicz (1985) attributed to the early infiltration of meteoric silica-bearing water. Siever's and Kantorowicz's arguments are plausible, but no Pleistocene or Holocene examples exist of such meteoric cementation except in silcretes,

91 which have characteristics different from the deposits they studied. Fluvial-deltaic sandstones contain less quartz cement than deeper-buried braided sandstones of the Travis Peak (Cretaceous) Formation of Texas (Dutton, 1987). Braided stream deposits have good lateral continuity and better transrnissivity and, thus, sequestered more quarz cement than shale-encased sandstones of the fluvial-deltaic facies. The environmental control of cement in these rocks was indirect.
The water volume problem

Evidence summarized above indicates that most silica for quartz cement is introduced to sand beds from external sources. Diffusion of silica from adjacent shale beds into sandstones undergoing cementation has been proposed (e.g., Fiichtbauer, 1967b; Jacka, 1970; Wood and Surdam, 1979; Lahann, 1980). It is difficult to understand how diffusion from immediately adjacent beds could supply more than a few percent of quartz cement, so most authors have favored the idea that silica was introduced from circulating groundwater. Calculations of the amount of water needed, either meteoric or briney basinal water, to introduce from 5 to 15% quartz cement yield extremely large volumes. This raises the question of the source of this water. Existing data (Fig. 1) suggest that most sandstones were cemented below the depth of most active meteoric groundwater flow by rising, cooling formation water saturated with silica (e.g., Siever, 1959, Land and Dutton, 1978; Boles and Franks, 1979; Mack, 1984; Leder and Park, 1986; Land et al., 1987). The amount of water required to introduce silica for quartz cement has been calculated for several quartz-rich sandstones. The assumption, for example, is that water at 100C that is saturated with silica rises to the depth where the temperature is 60 C and precipitates the supersaturated silica as quartz cement. Most calculations yield values in the range of 1 0 4 to 105 pore volumes (i.e., pre-cement pore volumes) of water necessary to introduce 5 to 15% cement in each cubic centimeter of rock

(von Englehardt; 1967; Land and Dutton, 1978; Bjorlykke, 1979, 1983; Mack, 1984; and Dutton and Land, 1988). This volume of water from shales undergoing compaction and clay mineral diagenesis is not available in most basins (Bjorlykke, 1979; Land and Dutton, 1979; Cathles, 1981; Land, 1984), unless special conditions exist of highly focussed rather than diffuse flow (e.g., Mack, 1984; Bodner, 1985; Pye and Krinsley, 1985; Prezbindowski and Pittman, 1987; Dutton, 1987; Pfeiffer, 1988). Recirculation of water through convective flow (Wood and Hewett, 1982; 1984) in thick basinal sequences is an attractive hypothesis to explain the volumes of water required to introduce quartz cement, as well as other cements, and the acids needed to produce secondary porosity. Convective flow of formation water in the thin sedimentary cover on oceanic crust has been modeled (Cathles, 1981) and called on to explain diagenetic reactions in the sediments (Lee and Klein, 1986). Fluid convection has been inferred to explain cement and secondary porosity patterns in passive-margin basins such as the Gabon (Cassan et al., 1981), and North Sea basins (Haszeldine et al., 1984), and is being evaluated as a diagenetic agent in the Gulf Coast basin (Land, 1984; Land et al., 1987). Convection is a fundamental part of the Leder and Park (1986) quartz cementation model. The Wood-Hewett (1984) model predicts that, within convection cells, dissolved silica should move from hot source zones, such as synclines, to cooler zones where silica is precipitated, such as anticlines. The model assumes homogeneous sands that are 10 to 100 m thick and porosities (20-30%) and permeabilities (1-10 darcies) that are found in few sandstones, however. Moreover, the segregation of calcite cement in synclines and quartz cement in anticlines as predicted by the model for folded strata has not been documented. Still, the water-shortage problem demands additional tests of the fluid convection hypothesis in thick sedimentary piles.

92

A few authors have supported the hypothesis that silica was introduced from descending meteoric water supersaturated only with respect to quartz (Van Hise, 1904; Siever, 1959; Blatt, 1979; Leder and Park, 1986). In the meteoric flow model analyzed by Blatt (1979), significant amounts of quartz cement could be formed in realistic geologic time only if groundwater with an assumed silica concentration of 33 p p m SiO 2 first descended and then rose vertically during the cementation process. Blatt's model is probably applicable to cratonic and some foreland basin settings, where advective flow of meteoric water likely dominates diagenesis (Bjorlykke, 1983; Garven and Freeze, 1984a, b). However, independent evidence from petrographic data, isotopic data or fluid inclusions is lacking which suggests that significant cementation by quartz took place at shallow burial depths. The importance of thermobaric water (term of Galloway, 1984) as a source of quartz cement remains uncertain also. Water derived from mineral dehydration reactions (Fyfe, 1973) during regional metamorphism of deeply buried sediments potentially is an important component of deep, advecting formation water. Such hydrothermal water was suggested recently as a source of the abundant quartz cement in Mesozoic sandstones of the Gulf Coast basin (Land et al., 1987; Sharp et al., 1989). Invasion of Mesozoic sandstones, but not Tertiary sandstones, with silica-rich hydrothermal fluids may explain the much greater amount of quartz cement in the Mesozoic sandstones (Table VI). Our understanding of sedimentary basin hydrology is too rudimentary at this time to help constrain hypothesis of formation water flow that produces quartz cementation. However, studies of quartz and carbonate cement distribution in anticlinal structures should provide a means of testing the Wood-Hewett convection model, and quartz cementation patterns around faults should test whether faults commonly serve as major fluid conduits as many workers suggest (e.g., Jones,

1975; Bodner, 1985; Evans, 1987). Studies cited previously show that focussed flow of formation water is potentially important in explaining inhomogeneous distribution of quartz cement in a number of formations.
SOURCES OF QUARTZ CEMENT

Many different sources of silica have been proposed as the source of quartz cement in sandstones. Reviews treating this particular problem have been presented by Travis (1963), Pittman (1972, 1979), and Jonas and McBride (1977). The following discussion is an update from the latter source. Proposed sources listed in chronological order with the first advocate are shown in Table I. Most authors recognize the likelihood that silica in quartz cement was derived from more than one source, but many interpret a particular source for a given formation to have been dominant. Pressure solution of quartz grains in sand and smectite-illite conversion in shale are the most often recently cited sources of silica for moderately to deeply buried sandstones, and Pittman (1979) listed the replacement of silicates by carbonate minerals and decomposition of feldspars as two additional important sources. In addition, numerous workers advocate meteoric water as a major cementing agent. The source of silica in meteoric water is from weathering or from dissolution of silicates in the shallow subsurface. However, the lack of a means of chemically identifying the source of silica in quartz cement will permit considerable debate to continue. Specific silica sources are reviewed below. (1) Pressure solution. Since Waldschmidt (1941), many authors have cited pressure solution of detrital quartz and other silicates at grain contacts in sandstones as an important source of silica for quartz cement, and some authors consider it to be the most important source (e.g., Waldschmidt, 1941; Pittman, 1972; Fiichtbauer, 1974a, p. 141), especially in sands that have undergone significant burial. Other pressure solution

93
TABLE VII Silica budget of pressure solution versus quartz cement Formation Bethel Muddy Knox Tensleep Grimsby Tuscarora St. Peter Tensleep Cisco Travis Peak Norphlet Various Simpson and St. Peter Hartshorne Nugget Age Mississippian Cretaceous Paleozoic Cambro-Ord. Pennsylvanian Silurian Silurian Ordovician Pennsylvanian Pennsylvanian Cretaceous Jurassic Paleozoic Ordovician Pennsylvanian Jurassic Location Illinois, USA Colorado, USA Appalachians, USA Virginia, USA Wyoming, USA Appalachians, USA Appalachians, USA Mid-continent, USA Wyoming, USA Texas, USA Texas, USA Mississippi, USA Appalachians, USA Mid-continent, USA Arkoma Basin, USA Colorado Plateau, USA Reference Pye (1944) Goldstein (1948) Heald (1950) Dietrich (1953) Todd (1963) Martini (1972) Sibley and Blatt (1974) Odom et al. (1979) Mankiewicz and Steidtman (1979) Mack (1984) Dutton (1986), (1987) McBride et al. (1987) Lowry (1956) Heald (1956) Houseknecht (1984) James et al. (1986)

Q u a r t z c e m e n t > p r e s s u r e solution

Pressure solution > quartz c e m e n t

sources that were suggested include stylolite seams in sandstones (Heald, 1955; Dutton, 1986), quartz grains in clayey siltstones (Fiichtbauer, 1967b), quartz grains in shales (Fiichtbauer, 1974a), and quartz and other silicates in sandstones and shales undergoing burial metamorphism (Land et al., 1987). Pressure solution of quartz and other silicates takes place in conglomerates and cherts (e.g., McBride and Thomson, 1970), but these rocks are important only locally because of their limited abundance. Questions arise whether the volume of silica present as cement could have been generated internally within a bed or formation by pressure solution, and about the timing of pressure solution in a bed relative to cementation. However, numerous authors report considerably more quartz cement in a formation than can be accounted for by pressure solution within the same formation (Table VII). Few of these studies were quantitative, few workers used cathodoluminescence, and, as reported by Pittman (1979), Sibley and Blatt's study ignored the finer-graind sandstones, which

generally show more severe pressure solution than coarser-grained beds. Grain pressure solution is size-dependent (Renton et al., 1969; Stephan, 1970; Pittman, 1979; Houseknecht, 1984; James and Porter, 1985; Porter and James, 1986). The latter authors reported pressure solution in very fine, fine and medium sand respectively to be 3.3, 2.5, and 2 times more extensive than in coarse sand as measured by the ratio of overlap q u a r t z / framework quartz. In support of a local pressure solution source of silica are the studies by Lowry (1956) and Heald (1956), who reported an equivalence of quartz cement and silica loss by pressure solution in some Appalachian Paleozoic sandstones and mid-continent Ordovician sandstones, respectively; Houseknecht (1984), who found a mass balance of quartz in the Hartshorne Sandstone (Penn.) in the Arkoma basin if formation water carried the silica 240 km laterally, and James et al. (1986), who reported the Nugget (Jurassic) to be an exporter of silica if the silica released at stylolites is added to that released by grain

94 TABLE VIII Depth of initiation of pressure solution Depth (m) 600 1000-1200 (chiefly 2000-4000 rn) > 2000 > 2000 - 1500 present depth, but > 1 km lost to erosion 2000 Formation Agua Grande Rotliegendes Age Cretaceous Permian Tertiary Eocene Cretaceous Triassic Reference Netto (1974) Ffichtbauer (1967b) Fothergill (1955) McBride (unpublished) Taylor (1950) Stephan (1970)

Wilcox (four different) Buntsandstone

pressure solution. Dutton (1986, pp. 39-40), McBride et al. (1987), McBride (1987a) and Houseknecht (1988) observed, however, that most quartz cementation in the sandstones that they studied took place before pressure solution began in the same formations. Thus, silica released by pressure solution in these two formations could not have been a major local source of silica for cement. Quantitative data on the depth at which pressure solution begins in siliciclastics are sparse and inconsistent (Table VIII). The shallowest depth reported for grain pressure solution in sandstones is 600 m in Brazil (Netto, 1974), but must reports indicate that burial of greater than 1500 m is needed to initiate significant pressure solution. Pressure solution seems to be relatively rare in Tertiary and Quaternary sandstones except at great burial depth. Time, temperature and overburden load are obvious variables that need to be evaluated (cf. Leder and Park, 1986). Experiments suggest that absolute pore pressure is more important than differential pressure (of fluid versus grains) in promoting pressure solution (Sprnnt and Nur, 1976). The increase in pressure solution of Lower Triassic Buntsandstone in proximity to a Late Mesozoic intrusive in West Germany shows the importance of temperature in pressure solution (Stephan, 1970), a point re-emphasized in a study of the Hartshorne Sandstone in the Arkoma basin (Houseknecht, 1984).

The studies summarized above concern pressure solution only in sandstones. Fiichtbauer (1967b) observed highly sutured quartz grains in Permian argillaceous siltstones and believed he could account for the amount of silica lost by pressure solution as cement in immediately overlying sandstones. He attributed strong pressure solution in the siltstones to the influence of clay in contact with quartz grains. The role of clay minerals in promoting pressure solution in terrigenous rocks has been emphasized by numerous authors including Thomson (1959) and Weyl (1959). A strong suspicion exists that pressure solution of quartz and possibly other silicate grains in shales may release considerable silica for quartz cement, because many quartz grains in shales have elongated, irregular shapes typical of those produced by pressure solution in sandstones and siltstones (Heald, 1955, fig. 3; FiJchtbauer, 1967b). Concretions in shales of various ages contain 10 to 50% more detrital quartz than host shale (Fiichtbauer, 1978), indicating a major dissolution loss of quartz from the shales. The role of pressure solution in the loss is uncertain. Pressure solution of deeply buried shales undergoing low-grade metamorphism is another potentially significant source of silica (Land et al., 1987). Large volume losses of quartz and other silicates have been documented during the development of slaty cleavage (e.g., Wright and Platt, 1982; Buetner and Charles, 1985; Henderson et al., 1986),

95 and the fate of this silica has not been determined. Vertical migration distances of several kilometers are necessary for this "metamorphic" source of silica to be introduced into most sandstone beds. Dewatering of shales during metamorphism is another source of water to transport silica for cement. (2) Conversion of clay minerals during burial diagenesis, chiefly smectite to illite. Burst (1959) was one of the first to document the extensive alterations that clay minerals undergo during burial diagenesis. He called attention to the alteration of smectite to illite and of kaolinite to either chlorite or illite. Siever (1962) wrote reactions for several transformations that release silica, and Towe (1962) suggested that the silica so released was a potentially large source of quartz cement. The following reactions have been proposed: smectite + K ~ illite + H 4 S i O 4 + cations (Ca, Mg, Na, Fe) transformation left the shales by diffusion, but accounted for only 10% of the cement present in some Mesozoic sandstones in Germany. The timing of clay diagenesis in a subsiding sediment pile must also be considered when evaluating the illitization of smectite as a major source of silica, because the clay reaction must precede or be coeval with quartz cementation if quartz cement is to be generated by this reaction. Land (1984, p. 51) reported that the temperature of quartz cementation of the Oligocene Frio Formation in the Texas Gulf Coast is cooler (and therefore took place shallower) than the temperature of smectite alteration, and Mack (1987) reported that Eocene Wilcox sandstones of the Texas Gulf Coast were cemented by quartz and later by calcite before smectite in Wilcox shales altered to illite. Thus, for these Tertiary sandstones either silica is released from smectite without the generation of much illite or there was another major source of silica. Fiichtbauer's (1974) study of concretions in shales documented a large dissolution loss of detrital quartz, although the role of pressure solution remains an issue. (3) Dissolution of detrital quartz grains in shale unrelated to pressure solution. Johnson (1920) argued that water compacted from shales was the main source of quartz cement in sandstones and implied that silica was derived from the dissolution of quartz silt grains and possibly other silicate minerals without the need for pressure solution. Because of their small size, silt grains were considered more likely to dissolve than sand without the influence of overburden pressure. Biederman (1962) and Sharma (1965) supported the non-pressure solution argument and suggested that quartz dust particles generated during eolian abrasion of sand-sized silt was the silica source. However, differences in solubility of sand-, silt- and claysized particles of quartz predicted by the Ostwald-Freundlich equation (Iler, 1979, p. 51), which considers both the particle curvature and interracial energy on solubility, are

(1)
smectite + K-feldspar ---, illite + quartz + chlorite (2)

Eq. 1 was proposed by Boles and Franks (1979) and eq. 2 by Hower et al. (1976). At least in the Gulf Coast, the abundance of smectite in shallow-buried Tertiary sequences suggests that reactions involving it are the most significant. The Boles and Franks equation predicts that shale 1 km thick can release sufficient silica to completely cement an overlying sandstone bed 100 m thick (Leder and Park, 1986). Little doubt exists that large volumes of silica are generated during diagenesis of clay minerals. Less certain is whether the silica so generated leaves the shales. Hower et al. (1976) and Yeh and Savin (1977) found evidence suggesting that the shales themselves acted as a sink for the newly released silica. Hower (1983), however, subsequently believed that some of the silica released from smectite did leave the shales. Fiichtbauer (1978) believed the silica generated by the

96 trivial. Applying the Ostwald-Freundlich equation to opaline silica particles indicates that, compared to the solubility of a flat plate, sand-size particles (100 # m diameter), silt-sized particles (10 #m) and clay-size particles (1 #m) are only 1.00002, 1.0002 and 1.002 times as soluble, respectively (Hurd and Birdwhistell, 1983; D. Hurd, pers. commun., 1987). Thus, if detrital quartz in shales follows the same grain-size solubility behavior as opaline silica and is a source of silica for cement, then pressure solution must have been a more important mechanism of dissolution than passive dissolution of very small grains. (4) Decomposition/kaoHnitization of feldspar, Sorby (1880) inferred that quartz cement was derived from the decomposition of feldspar, and Fothergill (1955) suggested specifically that kaolinitization was an important source of silica for cement. For every mole of K-feldspar altered to kaolinite, two moles of silica are released and made available for cement (Siever, 1957). Fothergill (1955) and Hawkins (1978) reported statistical correlations between kaolinitized feldspars and quartz overgrowths. However, many sandstones with abundant quartz cement have or had few feldspars, and most arkoses have more clay and carbonate cements than quartz cement (Table III). Silica derived from feldspar alteration does not appear to be a major source of quartz cement. In the past ten years or so it has become apparent that much detrital plagioclase and K-feldspar is lost by dissolution (e.g., McBride, 1979; Gold; 1984; Land, 1984; Land et al., 1987) rather than by replacement. Some of this silica certainly ends up in authigenic clays, but some of it may form quartz cement. However, the time of major feldspar dissolution also post-dates the time of major cementation in Tertiary sandstones of the Texas Gulf Coast (Land et al., 1987). Silica released by feldspar alteration may form the common but small amount of late-stage quartz cement that grows in secondary pores as tiny crystals that apparently lack overgrowth fabric (Fig. 6). (5) Sea water. Krynine (1941) believed that most quartz cement, at least in quartzarenites, was precipitated at the sediment-seawater interface. His argument was based on the observation that "orthoquarztites" have fairly loose grain packing, precluding compaction, and the absence of grain pressure solution, which Waldschmidt (1941) argued was the main source of cement. Amaral (1975, p. 256) suggested that sufficient quartz dissolved in seawater during the bed-load abrasion of detrital quartz grains that it was later precipitated below the sediment-water interface. Analyses of dissolved silica content of marine bottom waters and of near-surface sediment pore waters indicate that both commonly approach opal saturation and thus are supersaturated with respect to quartz (Hesse, 1986), but little evidence exists to suggest that significant quartz cementation begins at the sediment-water interface in the modern ocean. Failure of quartz to precipitate from m o d e m supersaturated surface waters has been attributed to kinetics of the precipitation reaction, but tiny quartz overgrowths have been precipitated experimentally in one year at 2 0 C from seawater saturated with quartz (Mackenzie and Gees, 1971). Four reports of quartz overgrowths forming in m o d e m subaqueous sands exist. Dapples (1967b, p. 325) reported small overgrowths on quartz-rich marine beach sands from New Jersey and from fresh-water Lake Michigan. However, sands from both areas contain recycled quartz grains, and the possibility that the grains with overgrowths are recycled cannot be dismissed. Baltzer and Le Ribault (1971) used the SEM to document the presence of microcrystalline and megaquartz overgrowths, some on a coating of opal, on quartz grains from tidal channels in New Caledonia. The authors attribute the precipitation of silica to the mixing of riverwater and seawater during Holocene tidal exchanges. Their short note is not clear on the sample control nor how they ruled out the possibility that the grams were not recycled. The precipitation of overgrowths, visible only

97 with the SEM, and chiefly in depressions on grains, is inferred to be an active process on the Labrador shelf (Fillon et al., 1978). Glacially produced quartz dust is the inferred local silica source for this cement. (6) Dissolution of opaline skeletal grains. Siliceous sponge spicules (Ordovician-Holocene), radiolaria (Cambrian-Holocene), diatoms (Triassic-Holocene), and silicoflagellates (Cretaceous-Holocene) secrete skeletal components composed of opal. Individuals of these taxa occur scattered in marine muds and they comprise more than 30% of certain deep-sea siliceous oozes. Siever (1957) suggested that these skeletal grains in shales are a possible significant source of cement silica in sandstones. Amorphous silica solubility is more than an order of magnitude greater than quartz (Siever, 1957), making opaline skeletal material an attractive potential source of cement silica. Hurst and Irwin (1982) and Pettijohn et al. (1987, p. 453) renewed the argument that early quartz cement in marine sands is derived from biogenic silica that dissolves shortly after deposition. Hurst and Irwin (1982) report the precipitation of quartz and dissolution of opal take place at similar rates: 2.4.10 -6 g cm -3 yr -1 versus 3.6-10 - 6 g c m - 3 yr -1. Pettijohn et al. (1987) suggested that biogenic silica dissolved in marine sediments and subsequently precipitated as quartz cement in sands when the rocks were uplifted and exposed to artesian groundwater. Diatoms and radiolaria are generally abundant in shales associated with bedded cherts and diatomites (e.g., Folk and McBride, 1978), but there is little quantitative data on the abundance of opaline skeletal grains in the average prodelta or shelf shale. However, an example of the importance of biogenic silica as a major local source of silica for cement in chert beds is seen in Jurassic bedded cherts of Italy (McBride and Folk, 1979). Siliceous nodules in shale interbeds of the Jurassic bedded chert contain 70% radiolaria, whereas the host shale beds contain only 10% radiolaria. This relation suggests that large-scale dissolution of radiolaria in shales was a major source of silica cement for adjacent chert beds and that preservation of original radiolarian abundance was rare. Scouring rushes (Equisetum) and ferns (Pteris), grasses, and some other terrigenous plants contain opaline particles (phytoliths) within their stems or leaves. The process of silica extraction from rocks by plants was investigated by Lovering and Engel (1967). Cayeux (1929, p. 249) made the novel suggestion that beds rich in deposits of the former two plants were possibly important local sources of silica for quartz cement in sandstones. Wind-blown phytoliths are the presumed source of silica in "turtle skin" crusts in modem desert dunes in Australia (Folk, 1978), and a potential source of silica in some silcretes (McBride, 1988). In the absence of a means of tagging silica molecules of specific parentage, the biogenic source of silica is difficult to assess. (7) Dissolution of quartz grains in sandstone unrelated to pressure solution. Goldstein (1948) first suggested that silica for quartz cement was derived from the dissolution of finer detrital quartz grains in sandstones without pressure solution, and other authors expressed variations on this theme. Dapples (1967) suggested that detrital quartz and chert grains were susceptible to dissolution, and Riezebos (1974) suggested that dislocation zones within detrital quartz would be more soluble than normal grain surfaces. Pye and Krinsley (1985) believed hot, alkaline fluids derived from underlying evaporites were effective in dissolving entire quartz grains in the Permian Rotliegendes Formation of the North Sea. They emphasized the dissolved quartz as a means of generating secondary pores, not as a source of quartz cement. Cassan et al. (1981) and Leder and Park (1986) suggested that quartz cement was derived internally from dissolution of quartz grains, but did not specify particular grain types. Waugh (1970) suggested that eolian sands generate 2 /~m chips from quartz grain edges and 50 ~m chips from grain faces, and that these abrasion dust particles are susceptible to being

98

dissolved in desert moisture and groundwater to provide a local source of quartz cement. These hypotheses also are difficult to evaluate, and require unusual circumstances. (8) Meteoric groundwater. Van Hise (1904) suggested that meteoric groundwater dissolves silica from silicates in the zone of weathering and moves downdip to precipitate quartz overgrowths. Siever (1959, p. 74) revived the idea with his suggestion that fluvial sands may undergo early cementation by quartz when riverwater and shallow groundwater mix. Several subsequent workers reported small amounts of quartz cement in sandstones invaded by meteoric waters under non-silcrete forming conditions (e.g., Walker, 1967; Stanley and Benson, 1979; Kantorowicz, 1985; Molenaar, 1986; Dutta and Suttner, 1986). Molenaar (1986) and Dutta and Suttner (1986) studied Paleozoic red beds and Walker (1967) studied Tertiary red beds; all interpreted small amounts of quartz cement to have been precipitated by meteoric groundwater. Kantorowicz (1985) reported early quartz cement in the Ravenscar Group (Jurassic) to be dominant in the fluvial facies, and he favored a meteoric source where silica was derived from silicate mineral dissolution as envisaged by Van Hise. A contrasting hypothesis is that of Alimen (1936), who suggested that quartz cement in the Oligocene Fontainebleau Sandstone in France formed as the result of evaporation of groundwater connected to interdune lagoons. Blatt (1966, 1979) has argued that only meteoric water possesses sufficient dissolved silica and flux to produce abundant quartz cement that is present in many quartzarenites. The literature suggests, however, that the amount of quartz cement precipitated by meteoric water in most ancient formations is quite small, and that only in silcrete sandstones is much quartz cement present. No direct evidence exists that sand aquifers today are being cemented by quartz, although studies of the Carrizo-Wilcox aquifer in Texas show that silicic acid is higher in groundwater of recharge areas than in discharge areas (Fogg

and Kreitler, 1982; Kaiser and Ambrose, 1986; Macpherson, 1986). The silica content of water in recharge areas is attributed to the dissolution of feldspar or other silicates. The authors infer that either quartz or clays are being precipitated along the flow path. Petrographic studies have not yet identified the fate of the silica in these shallow subsurface samples. However, Blatt's thesis has not been adequately tested. Evidence against the importance of meteoric water as a cementing agent cited above is not from intracratonic or foreland basins typical of the occurrence of quartz-cemented quartzose sandstones, and quartz cement in the subject sandstones has not been studied using fluid inclusions or oxygen isotopes. As reported earlier, small amounts of quartz cement are reported on quartz grains fractured by glacial transport and subject to emersion in glacial meltwater (Whalley and Krinsley, 1974; Fillon et al., 1978). Most sandstones for which data are available have been cemented at depths greater than 2000 m (Table II; Fig. 1). Nevertheless, isotopic evidence indicates that quartz cement in many sandstones is precipitated by formation waters with a strong meteoric component (e.g., Fisher, 1982; McBride et al., 1987; Dutton and Land, 1988). This water reaches depths of from 2 to 4 km in the Gulf Coast basin and may be involved in thermal convection (see a previous section). Precipitation of quartz from formation waters with a strong meteoric signature apparently occurs as the waters cool during upward migration. (9) Replacement of quartz and silicate grains by carbonate. Almost every sandstone that contains some chemically p r e c i p i t a t e d carbonate minerals displays evidence of replacement of feldspar or other silicate grains, and in some sandstones the margins of detrital quartz grains themselves. The importance of silica-carbonate replacements and their reversibility was reported by Walker (1960). M a n y p e t r o g r a p h e r s report significant amounts of quartz being replaced by carbonate cement based upon finding embayed

99 contacts where carbonate cement borders detrital quartz grains. This is unreliable evidence of replacement (Glover, 1963) and leads, I think, to an overestimation of the importance of this process. For example, thin sections of the Maxon Formation (Cretaceous) show irregular contacts of calcite cement against quartz grains, but few quartz grains show any evidence of corrosion when the cement is removed with acid (McBride, 1987). The only quantitative data that have been presented to document the volume of silica lost in a formation by etching of carbonate is that of Burley and Kantorowicz (1986), who demonstrated that the surfaces of detrital quartz grains were embayed and corroded by calcite cement to a significant degree. More quantitative data are needed to assess the importance of this mechanism. (10) Amorphous alumino-silicates that are detrital. Fisher (1982) suggested that detrital amorphous aluminosilicates transported with clays or on sand grains were a possible source of silica for quartz cement in the Eocene Wilcox Formation in the Texas Gulf Coast. However, 15 Pliocene/Pleistocene Gulf Coast mudstone samples from depths of 800-4900 m were analysed for amorphous silica and aluminosilicates but contained only insignificant amounts of material (Land et al., 1987). The importance of Fisher's proposed source needs further testing. (11) Silica released from the hydrolysis of volcanic glass. Zen (1959) observed that silica can be released during hydrolysis of volcanic glass, Swineford and Franks (1959) suggested that silica from this source formed opal cement in near-surface sands of the Pliocene/ Miocene Ogallala Formation in Kansas, and Fiichtbauer (1974a, p. 140) suggested glass shards might be a potential source of quartz cement in sandstones. However, silica derived from glass at surface conditions precipitates at rates and from solutions of sufficient ionic strength to form chiefly opal and microcrystalline quartz (cf. Millot et al., 1963), whereas silica derived from glass in more deeply buried sands reacts with aluminium and other metal cations to form zeolite, clay minerals and perhaps albite (Surdam and Boles, 1979). No well documented example exists of a highly quartz-cemented sandstone having derived most of its silica from volcanic glass. (12) Silica dissolved from detrital quartz grains in coastal sands as organic complexes (Evans, 1965). Evans (1964) partly cemented quartz sand with silica in 7 days using 0.2 M adenosine triphosphate at room temperature. He later suggested (Evans, 1965, p. 19) that nucleic, alginic, and amino acids released by organisms, seaweeds and rotting plants in coastal environments might be capable of dissolving sufficient quartz from sand grains and releasing it soon afterward to cement marine sands. Bennett and Siegel (1987) documented dissolution of quartz and other silicates by oxidized crude oil spilled from a ruptured pipeline. Oxidation of crude oil generated various aromatic, hydroxy- and ketoacids that invaded the local groundwater system. The latter authors use this example to suggest that these organic acids may be involved in the cementation of sands at neutral pH and near-surface conditions. However, the rarity of quartz cement in modern marine sands indicates that cementation by organic acids is not rapid and probably not important volumetrically. Organic acids are effective in dissolving rock-forming silicate minerals, especially feldspars (Huang and Keller, 1970; Surdam et al., 1984), and the silica released that does not end up in clays may be available for quartz cement. Mono- and difunctional carboxylic acids are most abundant at depths slightly deeper (i.e., hotter = 77 o to 120 o C; Surdam et al., 1984, fig. 19) than the depths of most cementation by quartz, but they may contribute to the cementation process. The volume of such organic acids appears to be too small for them to be important cementing agents, and increased silica (or alumina) concentrati,ms in organic acid-rich formation waters of the Gulf Coast are not observed (L.S. Land, oral commun., 1988). (13) Desorption of silica from marine clays later exposed on coastal plains (Siever, 1971). Siever observed that river-transported clays

100 sorb silica when entering the ocean, and he speculated that if the clays were subsequently exposed on a prograding/uplifting coastal plain, they should release their sorbed silica. There is no evidence at present to support this mechanism as being volumetrically important. (14) Silica released from the alteration of silicates other than quartz or feldspar. Van Hise (1904) proposed the decomposition of detrital silicates in general as the source of silica in meteoric water. This source of silica has been cited by numerous subsequent workers as possible important local sources in continental deposits and marginal marine deposits invaded by meteoric water (Walker, 1967; Fiichtbauer, 1974b; Kantorowicz, 1985; Dutta and Suttner, 1986). Robinson (1980) argued that even quartz can be a by-product of the weathering of quartz-free bedrock. Certainly the alteration of silicates is one source of silica in meteoric water. There is little evidence to indicate that meteoric water precipitates much quartz cement as noted earlier. (15) Silica introduced from hypersaline SILCRETES Silcretes are silica-indurated products of surface and near-surface diagenesis (Summerfield, 1983a, 1986). Terms roughly equivalent to silcrete include grey billy, billy, porcellanite (Australia), meulirre (France), and sarsen and puddingstone (Britain). The term ganister is used in Britain for a quartz-cemented sandstone that formed as a paleosol and which is suitable for use as raw material in the production of refractories (Percival, 1983). Most silcretes are known only from Cenozoic rocks, but this is probably a problem of recognizing them in older strata. They form both by the passive cementation of sands and gravels, locally producing non-porous sandstones with normal quartz overgrowths or with chert cement, and by the replacement of soils, producing rocks resembling cherts and chert breccias (Smale, 1973; Watts, 1978; Ambrose and Flint, 1981, Summerfield, 1983a, b; 1986). Silcretes form both at weathering profiles, where they conform to old topographic surfaces and show pedogenic textures and ferro-titanium oxides typical of soils, or at stable water tables, where they do not conform with topographic surfaces and generally show features typical of passive cementation. Both types of silcretes can have the same texture of quartz overgrowths as sandstones cemented at much greater depths (Summerfield, 1983a; Thiry et al., 1988). The authigenic silica in both pedogenic silcretes and groundwater silcretes (terms of Milnes and Thiry, 1986) can be composed of the spectrum of silica polymorphs (opal, microcrystalline quartz, and megaquartz) cited in the introduction to this review. Soil silcrete sandstones are more commonly cemented by opal and microcrystalline quartz, whereas water-table silcretes tend to have normal syntaxial quartz overgrowths. As mentioned earlier, the polymorphic form of silica apparently is controlled largely by the degree of supersaturation of silica and the purity of the groundwater and sediments (Millot, 1960; Thiry and Millot, 1987). The conditions under which silcretes form

brines derived from adjacent formations (Odom et al., 1979). This hypothesis was suggested to
explain a small amount of quartz cement in the St. Peter Sandstone (Ordovician), a stratigraphic unit which is not in close proximity to shale beds and which shows trivial pressure solution in the northern Illinois basin. Whereas some surface brines with continental affinities (alkali lakes) are alkaline and capable of carrying significant silica in solution (Peterson and v o n d e r Borch, 1965; Eugster, 1969), marine brines quickly turn acid, and, thus, are not likely to be major sources of silica. In fact the reduced activity of water in Cl-rich brines reduces the solubility of silica considerably. (16) Silica derived from glacial abrasion dust (Whalley and Krinsley, 1974). Micrometer-size particles of quartz produced by glacial abrasion of quartz grains were suggested as local sources of quartz cement, also of micrometer size, in glacial and pro-glacial deposits (Whalley and Krinsley, 1974) and marine environments (Fillon et al., 1978).

101

are known sketchily, but evidence suggests they form chiefly under a r i d / s e m i a r i d climates, but can form under deep weathering in humid, tropical climates also (Summerfield, 1983a; 1986). The sources of silica for silcretes remain largely conjectural, but the concensus is that silica is derived from weathering reactions of silicate minerals, including clays, and dissolution of quartz in bedrock and alluvium. The role of pH and organic acids in promoting silica dissolution also remains debatable. A distinctive genetic type of silcrete is that cemented largely by opal and which is interbedded with tuffaceous strata. Examples are the Miocene/Pliocene Ogallala Formation of the Great Plains states of the U.S. (Swineford and Franks, 1959) and the Carlos Member of the Jackson Formation (Eocene) of the Texas coastal plain. The bulk of the silica in these silcretes were derived from volcanic glass in adjacent tuff beds. Silica-bearing plants (Equisetum, hackberry endocarps, and diatoms) are also suggested contributors to the opal cement of the Ogallala. Water-table silcretes completely cemented by quartz can form in as little as 200,000 years (Thiry et al., 1988), but the time necessary to form pedogenic silcretes is unknown. Reviews of silcretes are given by Wopfner (1978) and Summerfield (1983a, b).
SUMMARY

Quartz cement as syntaxial overgrowths is one of the two most abundant cements in sandstones. Quartz cement averages approximately 12 vol.% in quartzarenites, but less than 3% in arkoses and litharenites; in only a few sandstones does quartz cement fill all pores. Arkoses and litharenites undergo more rapid burial and loss of porosity than quartzarenites. They contain many unstable feldspars and rock fragments whose hydrolysis releases A1 and other cations that tend to lock up silica in authigenic clay minerals and other silicates. The amount of quartz cement in sandstones of all clan types can be highly

variable at the hand specimen and formation scale. The main factors that control the amount of quartz cement in a sandstone are framework composition, residence time in the silica mobility window, and fluid composition, flow path and flow volume. These variables are controlled by tectonics: the type of sedimentary basin and composition and relief of the source area. Clay and other mineral coatings on detrital quartz grains and entrapment of hydrocarbons in pores retard or prevent cementation by quartz, whereas highly permeable sands that serve as major fluid conduits tend to sequester the greatest amounts of quartz cement. Quartz cement commonly is selective to coarser or finer grained beds or laminae, but the reasons are debatable. Our understanding of quartz cementation is biased by the abundance of data from the Gulf Coast basin and North Sea basin, both actively subsiding (or recently so) deep basins on passive margins. In these and similar basins most quartz cement is precipitated by rising, cooling formation water at burial depths of several kilometers where temperatures are from 60 o to 100 C. This conclusion is based on the distribution of quartz in basins where thermal gradients are known, on fluid inclusion data, and on oxygen isotopic data from quartz overgrowths. Cementation proceeds over millions of years, often under changing fluid compositions and temperatures. The similarity of diagenetic sequences in sandstones of different composition and ages is interpreted to be the result of the normal time-temperature-dependent maturation of sediments, organic matter and pore fluids during burial in a sedimentary basin. Silica that forms overgrowths is released by one or more temperature-time-dependent diagenetic processes in the sandstones themselves, in associated shales, or in more deeply buried siliciclastic rocks that undergo metamorphism. The relative time of cementation by quartz, then, is not a chance event but one that is predictable. The amount of quartz cement is apparently controlled by the amount

102

of time a sandstone resides within the silica mobility window or is permeated by fluids derived from within the window. Factors that influence the plumbing system, such as sandstone geometry, compartments developed by faults (e.g., North Sea) or overpressure/hydropressure boundaries (e.g., Gulf Coast), can play a major role in influencing the cementation pattern. Sandstones with more than 10% imported quartz cement pose special problems of fluid flux and silica transport. If silica is transported as H4SiO4, convective recycling of formation water seems to be essential to explain the volume of cement found in many sandstones. Precipitation from single-cycle, upward-migrating formation water is adequate to provide the volume of cement only if significant volumes of silica are transported in complexes that have not yet been identified or if basin-derived formation water flow has been highly focussed along certain conduits. Limited data suggest that tens of millions of years are required to introduce large volumes of quartz cement. Hydrologic modeling indicates that sandstones in intracratonic and foreland basins can be subjected to a much greater flux of meteoric water than compactional water. Thus, quartzarenites and other quartzose sandstones with abundant quartz cement in these types of basins were probably cemented by meteoric water, although evidence to confirm this hypothesis is lacking. Silica for quartz cement other than meteoric water comes from both shale and sandstone beds within the depositional basin, including possibly deeply buried rocks undergoing low-grade metamorphism, but the relative importance of potential sources remains controversial and likely differs for different formations. The most likely important silica sources within unmetamorphosed shales include clay transformation (chiefly illitization of smectite), dissolution/pressure solution of detrital quartz and possibly other silicate grains, and dissolution of opal skeletal grains; the most important sources of silica within unmetamorphosed sandstones includes pres-

sure solution of detrital quartz grains at grain contacts and at stylolites, feldspar alteration/ dissolution, and perhaps carbonate replacement of quartz and other silicate minerals. Silica released by pressure solution in a sandstone generally post-dates the episode of cementation by quartz; thus, this silica must migrate and cement shallower sandstones in the basin or escape altogether. Some quartzcemented sandstones are separated from potential silica source beds by a kilometer or more, requiring silica transport over long distances. Shallow meteoric water produces only small amounts of silica cement, generally less than 3% in most fluvial (and eolian?) sandstones, except in certain soils and at water tables in high-flux sand aquifers. Soil silcretes are commonly cemented by opal and microcrystalline quartz, whereas water-table silcretes are apparently likely to be cemented by normal syntaxial quartz overgrowths. Silica for silcrete cements and replacements comes from quartz, silicate minerals, and locally volcanic glass, in alluvium and bedrock. Strongly cemented silcretes can form in less than a million years. In the centennium since the work of Sorby we are still plagued by uncertainty about the relative importance of the sources of silica for quartz cement and an uncertainty of the hydrologic system(s) that accomplish cementation. Empirical schemes like that of Leder and Park (1986) permit a certain degree of accuracy in the prediction of the amount of quartz cement expected in quartz-rich sandstones. However, we are not yet able to make accurate predictions for most types of sandstones. More quantitative case studies are needed to be able to write a more inclusive set of rules that govern quartz cementation.
ACKNOWLEDGEMENTS

I greatly appreciate the editorial comments of manuscript drafts made by Bob Blodgett, Tim Diggs, Bob Folk, Dave Hurd, Lynton Land, Kitty Lou Milliken, and Ed Pittman. Marc Haws helped with word processing and

103

references. Karl Hoops provided some translations from the German. Journal reviewers R.L. Hay, David Houseknecht, and George deVries Klein suggested useful improvements in the manuscript. Dave Houseknecht's constructive comments are especially appreciated.
REFERENCES Aaron, J.M., 1969. Petrology and origin of the Hardyston Quartzite (Lower Cambrian) in eastern Pennsylvania and western New Jersey. In: S. Subitzky (Editor), Geology of Selected Areas in New Jersey and Eastern Pennsylvania and Guidebook of Excursions. Rutgers Univ. Press, New Brunswick, N.J., pp. 21-35. Alimen, A., 1936. l~tude sur le Stampien du Bassin de Paris, Soc. G~,ologique de France, Mem., V. 31,304 PP. Amaral, E.J., 1975. Textural Analysis of the St. Peter Sandstone, Southwestern Wisconsin. Master's Thesis, Univ. Cincinnati, 428 pp. Ambrose, G.J. and Flint, R.B., 1981. Regressive Miocene lake system and silicified strandline in northern South Australia: implications of regional stratigraphy and silcrete genesis. J. Geol. Soc. Austr., 28: 81-94. Anderson, D.W., and Picard, M.D., 1971. Quartz extinction in stiltsone. Geol. Soc. Am. Bull., 82: 181-185. Austin, G.S., 1974. Multiple overgrowths on detrital quartz sand grains in the Shakopee Formation (Lower Ordovician) of Minnesota. J. Sediment. Petrol., 44: 358-362. Bailey, J.L., 1983. Stratigraphy, environments of deposition, and petrography of the Cotton Valley Terryville Formation in eastern Texas. Master's Thesis, University of Texas at Austin, 228 pp. Baltzer, F. and Le Ribault, L., 1971. Ntogentse de quartz dans les bancs stdimentaires d'un delta tropical. Aspect des grains en microscopie 61ectronique et optique. C. R. Acad. Sci. Paris, 273: 1083-1086. Bennett, P. and Siegel, D.I., 1987. Increased solubility of quartz in water due to complexing by organic compounds. Nature, 326: 684-686. Berg, R.R. and Davies, D.K., 1968. Sandstone at Bell Creek Field, Montana. Am. Assoc. Petrol. Geol. Bull. 52: 1888-1899. Bethke, C.M., 1986. Hydrologic constraints on the genesis of the Upper Mississippi Valley mineral district from Illinois basin brines. Econ. Geol., 81: 233-249. Beutner, E.C. and Charles, E.G., 1985. Large volume loss during cleavage formation, Hamburg sequence, Pennsylvania. Geology, 13: 803-805.

Biederman, E.W., 1962. Distinction of shoreline environments in New Jersey. J. Sediment. Petrol., 32: 181-200. Bjorlykke, K., 1979. Cementation of sandstones--a discussion. Jour. Sediment. Petrol., 49: 1358-1359. Bjorlykke, K., 1983. Diagenetic reactions in sandstones: In: A. Parker and B.W. Sellwood (Editors), Sediment Diagenesis, Reidel, Dordrecht, pp. 169-213. Bjorlykke, K., Maim, O. and Everh~i, A., 1979. Diagenesis in the Mesozoic sandstones from Spitsbergern and the North Sea. Geol. Rundsch. 68: 1151-1171. Blanchard, P.E. and Sharp, J.M., Jr., 1985. Possible free convection in thick Gulf Coast sandstone sequences. In: C.L. McNulty and J.G. McPherson (Editors), Southwest Section. Am. Assoc. Petrol. Geol., Fort Worth Geol. Soc., pp. 6-12. Blatt, H., 1966. Diagenesis of sandstone: processes and problems. Wyoming Geological Assoc., 12th Annu. Conf., pp. 65A-70. Blatt, H., 1979. Diagenetic processes in sandstones. In: P.A. Scholle, and P.R. Schluger (Editors), Aspects of Diagenesis. Soc. Econ. Paieontol. Mineral. Spec. Publ., 26: 141-157. Bodner, D.P., 1985. Heat Variations Caused by Groundwater Flow in Growth Faults of the South Texas Gulf Coast Basin. Master's Thesis, University of Texas, Austin, Texas, 185 pp. Boggs, S., Jr., 1966. Petrology of the Minturn Formation, east-central Eagle County, Colorado. Am. Assoc. Petrol. Geol. Bull., 50: 1399-1422. Boles, J.R. and Franks, S.G., 1979. Clay diagenesis in Wilcox sandstones of southwest Texas: implications of smectite diagenesis on sandstone cementation. J. Sediment. Petrol., 49: 55-70. Burley, S.D. and Kantorowicz, J.D., 1986. Thin section and SEM textural criteria for the recognition of cement-dissolution porosity in sandstones. Sedimentology, 33: 587-604. Burst, J.F., 1959. Post diagenetic clay mineral environmental relationships in the Gulf Coast Eocene. Clays Clay Miner., 2: 327-341. Cassan, J-P., Garcia Palacios, M.C., Fritz, B. and Tardy, Y., 1981. Diagenesis of sandstone reservoirs as shown by petrographical and geochemical analysis of oilbearing formations in the Gabon Basin. Bull. Centres Rech. Explor. Prod. Elf-Aquitaine, 5: 113-135. Cathles, L.M., 1981. Fluid flow and genesis of hydrothermal ore deposits. In: B.J. Skinner (Editor), Economic Geology, 75th Ann. Volume, pp. 424-457. Cathles, L.M. and Smith, A.T., 1983. Thermal constraints on the formation of Mississippi valley-type lead-zinc deposits and their implications for episodic basin dewatering and deposit genesis. Econ. Geol., 78: 983-1002. Cayeux, L., 1929. Les Roches Stdimentaires de France: Roches Siliceuses. Masson, Paris, 250 pp.

104 Chang, H.K., 1983. Diagenesis and Mass Transfer in Cretaceous Sandstone-Shale Sequences, Offshore Brazil. Ph.D. diss., Northwestern University, Chicago, Ill., 337 pp. Chen, C.S. and Goodell, H.G., 1964. The petrology of Lower Pennsylvanian Sewanee Sandstone, Lookout Mountain, Alabama and Georgia. J. Sediment. Petrol., 34: 46-72. Crone, A.J., 1974. Experimental studies of mechanically-infiltrated clay matrix in sand (Abstr.). Geol. Soc. Am. Ann. Mtgs., 6: 701. Crook, K.A.W., 1960. Petrology of Parry Group, Upper Devonian-Lower Carboniferous, Tamworth-Nundle District, New South Wales. J. Sediment. Petrol., 30: 538-552. Dahl, W.M., 1984. Progressive-Burial Diagenesis in Lower Tuscaloosa Sandstones, Louisiana and Mississippi. Master's Thesis, University of New Orleans, 145 pp. Dapples, E.C., 1967a. Diagenesis of sandstone. In: G. Larsen and G.V. Chilingar (Editors), Diagenesis in Sediments. Developments in Sedimentology 8, Elsevier) Amsterdam, pp. 91-126. Dapples, E.C., 1967b. Silica as an agent in diagenesis. In: G. Larsen and G.V. Chilingar (Editors), Diagenesis in Sediments. Developments in Sedimentology 8, Elsevier, Amsterdam, pp. 323-342. Deneke, E. and Gunther, K., 1981. Petrology and arrangement of Tertiary graywacke and sandstone sequences of the northern Apennines. Sediment. Geol., 28: 189. Dickinson, W.R., 1985. Interpreting provenance relations from detrital modes of sandstones. In: G.G. Zuffa (Editor), Provenance of Arenites. Reidel, Dordrecht, 408 pp. Dietrich, R.V., 1953. Hexagonal chert prisms from the Knox dolomite, Va., J. Geol., 61: 65-68. Dixon, S.A., Summers, D.M. and Surdam, R.C., 1988. Diagenesis and preservation of porosity in the Norphlet formation (Upper Jurassic), southern Alabama. Am. Assoc. Petrol. Geol. Bull. In press. Dutta, P.K. 1986. Source of silicate and carbonate cements during deep burial diagenesis (Abstr.). Am. Assoc. Petrol. Geol. Bul., 70: 584. Dutta, P.K. and Suttner, L.J., 1986. Alluvial sandstone composition and paleoclimate, II. Authigenic mineralogy. J. Sediment. Petrol., 56: 346-358. Dutton, S.P., 1986. Diagenesis and Burial History of the Lower Cretaceous Travis Peak Formation, East Texas. Ph.D. diss., University of Texas, Austin, Texas, 165 pp. Dutton, S.P., 1987. Diagenesis and burial history of the Lower Cretaceous Travis Peak Formation, East Texas. Bureau of Econ. Geology Report of Investigations, No. 164, 58 pp. Dutton, S.P. and Land, L.S., 1985. Meteoric burial diagenesis of Pennsylvanian arkosic sandstones, southwestern Anadarko Basin, Texas. Am. Assoc. Petrol. Geol. Bull., 69: 22-38. Dutton, S.P. and Land, L.S., 1988. Diagenetic history of a well-cemented quartzarenite, Lower Cretaceous Travis Peak Formation, East Texas. Geol. Soc. Am. Bull., 100: 1271-1282. Ernst, W.G. and Blatt, H., 1964. Experimental study of quartz overgrowths and synthetic quartzites. J. Geol., 72: 461-470. Eugster, H.P., 1969. Inorganic bedded cherts from the Magadi area, Kenya. Contrib. Mineral. Petrol., 22: 1-31. Evans, R., 1987. Pathways of migration of oil and gas in the South Mississippi Salt Basin. Gulf Coast Assoc. Geol. Soc. Trans., 37: 75-86. Evans, W.D., 1964, The organic solubilization of minerals sediments. In: U. Colombo and G.D. Hobson (Editors), Advances in Organic Geochemistry. MacMillan, New York, N.Y., 488 pp. Evans, W.D., 1965. Facets of organic geochemistry. In: E.G. Hallsworth and D.V. Crawford (Editors), Experimental Pedology. Butterworths, London, pp. 14-28. Fielder, G.W., DiStefano, M.P. and Shearer, J.N., 1985. Depositional influences in sandstone diagenesis of the Lower Cretaceous Hosston Formation, Marion and Walthall Counties, Mississippi. Gulf Coast Assoc. Geol. Soc. Trans., 35: 367-372. Fillon, R.H., Ferguson, C. and Thomas, F., 1978. Cenozoic provenance and sediment cycling: Hamilton Bank, Labrador shelf. J. Sediment. Petrol., 48: 253-268. Fisher, R.S., 1982. Diagenetic History of Eocene Wilcox Sandstones and Associated Formation Waters, South-Central Texas. Ph.D. diss., University of Texas, Austin, Texas, 185 pp. Fisher, R.S. and Land, L.S., 1986. Diagenetic history of Eocene Wilcox sandstones, south-central Texas. Geochim. Cosmochim. Acta, 50: 551-561. Fogg, G.E. and Kreitler, C.W., 1982. Ground-water hydraulics and hydrochemical facies in Eocene aquifers of the East Texas Basin. Burea of Econ. Geology Report of Investigations, No. 127, University of Texas at Austin, 75 pp. Folk, R.L., 1968. Petrology of Sedimentary Rocks. Hemphill Publishing Co., Austin, Texas, 170 pp. Folk, R.L., 1978. Angularity and silica coatings of Simpson Desert sand grains, Northern Territory, Australia. J. Sediment. Petrol., 48: 611-624. Folk, R.L. and McBride, E.F., 1978. Origine pedogenica delle oficalciti ligerie: serpentiniti calcificate durante il Gurassico. Ofioliti, 3: 177-188. Fondeur, C., 1964. Etude p&rographique drtaillre d'un grrs a structure en feuillets. Rev. Inst. Fran~. Prtrole, 19: 901-920.

105 Fothergill, C.A., 1955. The cementation of oil reservoir sands and its origin. Proc. 4th World Petrol. C o n g . Sec. 1B, Paper 1, pp. 301-314. Fox, J.E., Lambert, P.W., Mast, R.F., Nuss, N.W. and Rein, R.D., 1975. Porosity variation in the Tensleep and its equivalent, the Weber Sandstone, western Wyoming: a log and petrographic analysis. In: D.W. Bolyard. (Editor), Symposium on Deep Drilling Frontiers in the Central Rocky Mountains. Rocky Mountain Assoc. of Geologists, Denver, Colo., pp. 185-216. Franks, P.C., 1969. Nature, origin, and significance of cone-in-cone structures in the Kiowa Formation (Early Cretaceous), north-central Kansas. J. Sediment. Petrol., 39: 1438-1454. Franks, S.G., 1980. Origin of porosity in deeply buried Tuscaloosa sandstones, False River Field, Louisiana (Abstr.) Gulf Coast Section, Soc. Econ. Paleontol. Mineral., First Annu. Res. Conf., p. 20. Franks, S.G. and Forester, R.W., 1984. Relatioships among secondary porosity, pore-fluid chemistry and carbon dioxide, Texas Gulf Coast. In: D.A. McDonald and R.C. Surdam (Editors), Clastic Diagenesis. Am. Assoc. Petrol. Geol. Mem. 37: 63-80. Friedman, M., 1954. Miocene orthoquartzite from New Jersey. J. Sediment. Petrol., 24: 235-241. FiJchtbauer, H., 1964. Sedimentpetrographische Untersuchungen in der ~ilteren Molasse n~Srdlich der Alpen. Eclogae Geol. Helv., 57: 157-298. Fiichtbauer, H., 1967a. Der Einflug des Ablagerungsmilieus auf die Sandstein-Diagenese im Mittleren Buntsandstein. Sediment. Geol., 1: 159-179. Fiichtbauer, H., 1967b. Influence of different types of diagenesis on sandstone porosity. 7th World Petrol. Congr., 2: 353-369. Fiichtbauer, H., 1974a. Sediments and Sedimentary Rocks, 1. Wiley, New York, N.Y., 464 pp. Fiichtbauer, H., 1974b. Zur Diageneses fluviatiler sandsteine. Geol. Rundsch., 63: 904-925. Fiichtbauer, H., 1978. Zur Herkunft des Quarzzements. Absch~itzung der QuartzaufliSsung in Silt- und Sandsteinen. Geol. Rundsch., 67: 991-1008. Fyfe, W.A., 1973. Dehydration reactions. Am. Assoc. Petrol. Geol. Bull., 57: 190-197. Galloway, W.E., 1984. Hydrologic regimes of sandstone diagenesis. In: D.A. McDonald and R.C. Surdam (Editors), Clastic Diagenesis. Am. Assos. Petrol. Geol. Mem., 37: 3-14. Garven, G. and Freeze, R.A., 1984a. Theoretical analysis of the role of groundwater flow in the genesis of stratabound ore deposits, 1. Mathematical and numerical model. Am. J. Sci., 284: 1085-1124. Garven, G. and Freeze, R.A., 1984b. Theoretical analysis of the role of groundwater flow in the genesis of stratabound ore deposits, 2. Quantitative results. Am. J. Sci., 284: 1125-1174. Ghosh, S.K. and Aguado, B., 1985. Diagenetic evolution of a clastic reservoir, Maracaibo Area, Venezuela (Abstr.). Soc. Econ. Paleontol. Mineral. Midyear Mtg., 2: 34. Glover, J.E., 1963. Studies in the diagenesis of some Western Australian sedimentary rocks. J. R. Soc. Aust., 46: 33-56. Gold, P.B., 1984. Diagenesis of Middle and Upper Miocene Sandstones, Lousiana Gulf Coast. Master's Thesis, University of Texas, Austin, Texas, 160 pp. Goldstein, A., Jr., 1948. Cementation of Dakota sandstone of the Colorado Front Range. J. Sediment Petrol., 18: 108-125. Grutzeck, M.W., 1986. St. Peter Sandstone: a closer look. J. Sediment. Petrol., 56: 669-673. Haszeldine, R.S., Samson, I.M. and Cornford, C., 1984. Quartz diagenesis and convective fluid movement; Beatrice Oilfield, UK North Sea. In: D.J. Morgan et al. (Editors), Patterns of Mineral Diagenesis on the Northwest European Continental Shelf and their Relations to Facies and Hydrocarbon Accumulation. Clay Miner., 19: 391-402. Hawkins, P.J., 1978. Relationship between diagenesis porosity reduction and oil emplacement in late Carboniferous sandstone reservoirs, Bothamsall Oilfield, E. Midlands. J. Geol. Soc. London, 135: 7-24. Heald, M.T., 1950. Authigenesis in West Virginia sandstones. J. Geol., 58: 624-633. Heald, M.T., 1953. Significance of stylohtes in sandstones (Abstr.). Geol. Soc. Am. Bull., 64: 1432. Heald, M.T., 1955. Stylolites in sandstones. J. Geol., 63: 101-114. Heald, M.T., 1956. Cementation of Simpson and St. Peter sandstones in parts of Oklahoma, Arkansas, and Missouri. J. Geol., 64: 16-30. Heald, M.T., 1956. Cementation of Triassic arkoses in Connecticut and Massachusetts. Geol. Soc. Am. Bull., 67: 1133-1154. Heald, M.T. and Anderegg, R.C., 1960. Differential cementation in the Tuscarora sandstone. J. Sediment. Petrol., 48: 568-577. Heald, M.T. and Renton, J.J., 1966. Experimental study of sandstone cementation. J. Sediment. Petrol., 36: 972-991. Heald, M.T. and Larese, R.E., 1974. Influence of coatings on quartz cementation. J. Sediment. Petrol., 44: 1269-1274. Heling, D., 1965. Zur Petrographic des Schilfsandsteins. Beitr. Miner. Petrogr., 11: 272-296. Henderson, J.R., Wright, T.O. and Henderson, M.N., 1986. A history of cleavage and folding: an example from the Goldenville Formation, Nova Scotia. Geol. Soc. Am. Bull., 97: 1354-1366. Henry, D.J., Toney, J.B., Suchecki, R.K. and Bloch, S., 1986. Development of quartz overgrowths and pressure solution in quartz sandstones: evidence from

106 cathodoluminescence backscattered electron imaging and trace element analysis on the electron microprobe (Abstr.) Geol. Soc. Am. Abstr. Progr., 18: 635. Hesse, R., 1986. Early diagenetic pore water/sediment interaction: modern offshore basins. Geosci. Can., 13: 165-196. Hoholick, D.J., 1980. Porosity, Grain Fabric, Water Chemistry, Cement, and Depth of the St. Peter Sandstone in the Illinois Basin. Master's Thesis, University of Cincinnati, Cincinnati, Ohio, 72 pp. Hoholick, D.J., Metarko, T. and Potter, P.E., 1984. Regional variations of porosity and cement: St. Peter and Mount Simon sandstones in Illinois Basin. Am. Assoc. Petrol. Geol. Bull., 68: 753-764. Houseknecht, D.W., 1984. Influence of grain size and temperature on intergranular pressure solution, quartz cementation, and porosity in a quartzose sandstone. J. Sediment. Petrol., 54: 348-361. Houseknecht, D.W., 1988. Intergranular pressure solution in four quartzose sandstones. J. Sediment. Petrol., 58: 228-246. Hower, J., 1983. Clay mineral reactions in clastic diagenesis (Abstr.). Am. Assoc. Petrol. Geol. Bull., 67: 486. Hower, J., Eslinger, E.V., Hower, M.E. and Perry, E.A., 1976. Mechanism of burial metamorphism of argillaceous sediment, I. Mineralogical and chemical evidence. Geol. Soc. Am. Bull., 87: 725-737. Huang, W.H. and Keller, W.D., 1970. Dissolution of rock-forming silicate minerals in organic acids: simulated first stage weathering of fresh mineral surfaces. Am. Mineral., 55: 2076-2094. Hubert, J.F., 1960. Petrology of the Fountain and Lyons Formations, Front Range, Colorado. Q. Colo. School Min., 55 (1) 242 pp. Hubert, J.F. and Reed, A.A., 1978. Red-bed diagenesis in the East Berline Formation, Newark Group, Connecticut Valley. J. Sediment. Petrol., 48: 175-184. Hurd, D.C. and Birdwhistell, S., 1983. On producing a more general model for biogenic silica dissolution. Am. J. Sci., 283: 1-28. Hurst, A. and Irwin, H., 1982. Geological modelling of clay diagenesis in sandstones. Clay Miner., 17: 5-22. Iler, R.K., 1979. The Chemistry of Silica: Solubility, Colloid and Surface Properties and Biochemistry. Wiley, New York, N.Y., 787 pp. Jacka, A.D., 1970. Principles of cementation and porosity-occlusion in Upper Cretaceous sandstones, Rocky Mountain region. 22nd Annu. Field Conf., Wyoming Geol. Assoc. Guidebook, pp. 265-285. James, W.C., 1985. Early diagenesis, Atherton Formation (Quaternary): a guide for understanding early cement distribution and grain modifications in nonmarine deposits. J. Sediment. Petrol., 55: 135-146. James, W.C. and Oaks, R.Q., Jr., 1977. Petrology of the Kinnikinic Quartzite (Middle Ordovician), eastcentral Idaho. J. Sediment. Petrol., 47: 1491-1511. James, W.C. and Porter, E.W., 1985, Influence of pressure, salinity, temperature, and grain size on silica diagenesis and reservoir quality in quartzose sandstones (Abstr.) Am. Assoc. Petrol. Geol. Bull., 69: 269. James, W.C., Wilmar, G.C. and Davidson, B.G., 1986. Role of quartz type and grain size in silica diagenesis, Nugget Sandstone, south-central Wyoming, J. Sediment Petrol., 56: 657-662. Johnson, R.H., 1920. The cementation process in sandstones. Am. Assoc. Petrol. Geol. Bull. 4: 33-35. Jonas, E.C. and McBride, E.F., 1977. Diagenesis of sandstone and shale: application to exploration for hydrocarbons. Continuing Education Program Publ. 1, The University of Texas, Austin, Texas, 165 pp. Jones, P.H., 1975. Geothermal and hydrocarbon regimes, northern Gulf of Mexico basin. Proc. 1st Geopressured Geothermal Energy Conf., pp. 15-89. Kaiser, W.R. and Ambrose, M.L., 1986. Hydrochemical mapping in the Wilcox-Carrizo Aquifer, Sabine Uplift area. In: W.R. Kaiser et al. (Editors) Geology and Ground-Water Hydrology of Deep-Basin Lignite in the Wilcox Group of East Texas. Burea of Economic Geology, University of Texas at Austin, pp. 85-99. Kantorowicz, J.D., 1985. The petrology and diagenesis of Middle Jurassic clastic sediments, Ravenscar Group, Yorkshire. Sedimentology, 32: 833-853. Krynine, P.D., 1941. Petrographic studies in variation in cementing material in the Oriskany sand. Proc. 10th Penn. Mineral Industries Conf., Penn. State College, 33: 108-116. Krynine, P.D., 1946. Microscopic morphology of quartz types. Anals Secondo Congresso Pan Americano Engenharia de Minas e Geologia, 35: 35-49. Krynine, P.D., 1948. The megascopic study and field classification of sedimentary rocks. J. Geol., 56: 130-165. Krynine, P.D., 1950. Petrology, stratigraphy and origin of the Triassic sedimentary rocks of Connecticut. Connecticut State Geol. Nat. Hist. Surv. Bull., 73, 273 pp. Lahann, R.W., 1980. Smectite diagenesis and sandstone cement. The effect of reaction temperature. J. Sediment. Petrol., 50: 755-760. Land, L.S., 1984. Frio sandstone diagenesis, Texas Gulf Coast. A regional isotopic study. In: D.A. McDonald and R.C. Surdam (Editors), Clastic Diagenesis. Am. Assoc. Petrol. Geol. Mem., 37: 47-62. Land, L.S. and Dutton, S.P. 1978. Cementation of a Pennsylvanian deltaic sandstone: isotopic data. J. Sediment. Petrol., 48: 1167-1176. Land, L.S. and Dutton, S.P. 1979. Cementation of sand-

107 stone--reply. J. Sediment. Petrol., 49: 1359-1361. Land, L.S., K.L. Milliken, and McBride, E.F., 1987. Diagenetic evolution of cenozoic sandstones, Gulf of Mexico sedimentary basin. Sediment. Geol., 50: 195-225. Larese, R.E., Pittman, E.D. and Heald, M.T., 1984. Effect of diagenesis on porosity development, Tuscaloosa sandstone, Louisiana (Abstr.) Am. Assoc. Petrol. Geol. Bull., 68: 498. Leder, F. and Park, W.C., 1986. Porosity reduction in sandstone by quartz overgrowth. Am. Assoc. Petrol. Geol. Bull., 70: 1713-1728. Lee, Y.I. and Klein, G. de V., 1986. Diagenesis of sandstones in the back-arc basins of the western Pacific Ocean. Sedimentology, 33: 651-675. Le Ribault, L., 1971. Presence d'une pellicule de silice amorphe h la surface de cristaux de quartz des formations sableuses. C. R. Acad. Sci. Paris, 272: 1933-1936. Levandowski, D.W., Kaley, M.E., Silverman, S.R. and R.G. Smalley, 1973. Cementation in Lyons Sandstone and its role in oil accumulation, Denver Basin, Colorado. Am. Assoc. Petrol. Geol. Bull., 57: 2217-2244. Lindquist, S.J., 1976. Sandstone Diagenesis and Reservoir Quality of Frio Formation (Oligocene), South Texas. Master's Thesis, Univ. of Texas at Austin, 148 pp. Link, M.H., Squires, R.L. and Colburn, I.P., 1984. Slope and deep-sea fan facies and paleogeography of Upper Cretaceous Chatsworth Formation, Simi Hills, California. Am. Assoc. Petrol. Geol. Bull., 68: 850-873. Livingstone, D.A., 1963. Data on geochemistry: chemical composition of rivers and lakses. U.S. Geol. Surv. Prof. Pap., 440-G, 64 pp. Longrnan, M.W., 1976. Depositional History, Paleoecology, and Diagenesis of the Bromide Formation (Ordovician), Arbuckle Mountains, Oklahoma. Ph.D. diss., University of Texas, Austin, Texas, 311 pp. Lonoy, A. and Akelsen, J., 1985. Early versus late diagenesis in a deeply buried reservoir: Hill Field, North Sea (Abstr.). Soc. Econ. Paleontol. Miner. Annu. Midyear Mtg., 2: 56. Loucks, R.G., 1978. Late silica diagenesis in the subsurface lower Cretaceous Pearsall Foundation, South Texas. Geol. Soc. Am. Abstr. Progr. 10: 21. Loucks, R.G., Bebout, D.G. and Galloway, W.E., 1977. Relationship of porosity formation and preservation to sandstone consolidation history--Gulf Coast Lower Tertiary Frio Formation. Gulf Coast Assoc. Geol. Soc. Trans., 27: 109-120. Loucks, R.G., Dodge, M.M. and Galloway, W.E., 1986. Controls on porosity and permeability of hydrocarbon reservoirs in Lower Tertiary sandstones along the Texas Gulf Coast. Bureau of Economic Geology Report of Investigations No. 149, The University of Texas, Austin, Texas, 78 pp. Lovering, T.G. and Engel, C., 1967. Translocation of silica and other elements from rock into Equisetum and three grasses. U.S. Geol. Surv. Prof. Pap., 594-B, 16 pp. Lowry, W.D., 1956. Factors in loss of porosity by quartzose sandstones of Virginia. Am. Assoc. Petrol. Geol. Bull., 40: 489-500. Mack, L.E., 1984. Petrography and Diagenesis of a Submarine Fan Sandstone, Cisco Group (Pennsylvanian), Nolan Country, Texas. Master's Thesis, University of Texas, Austin, Texas, 254 pp. Mack, L.E., 1987. Origin of silica for quartz cement in Wilcox sandstone: Sr isotopic evidence (Abstr.). Soc. Econ. Paleontol. Miner. Annu. Midyear Mtg. Abstr., 4: 50. Mackenzie, F.T. and Gees, R., 1971. Quartz synthesis at Earth-surface conditions. Science, 173: 533-534. Macpherson, G.L., 1986. Hydrochemistry of the Carrizo sand and the Wilcox Group in east-central Texas. In: W.R. Kaiser et al. (Editors), Geology and Ground-Water Hydrology of Deep-Basin Lignite in the Wilcox Group of East Texas. Bureau of Economic Geology, University of Texas at Austin, pp. 100-114. Mankiewicz, D. and Steidtmann, J.R., 1979. Depositional environments and diagenesis of the Tensleep Sandstone, eastern Big Horn Basin, Wyoming. In: P.A. Scholle and P.R. Schluger (Editors), Aspects of Diagenesis. Soc. Econ. Paleontol. Miner. Spec. Publ. 26, Tulsa, Okla., pp. 319-336. Martini, I.P., 1972. Studies of microfabrics: an analysis of packing in Grimsby Sandstone (Silurian), Ontario and New York State: Stratigraphy and Sedimentology. Int. Geol. Congr., Montreal, Sec. G, pp. 415-425. Matter, A. and Ramseyer, K., 1985. Cathodoluminescence microscopy as a tool for provenance studies of sandstones. In: G.G. Zuffa (Editor), Provenance of Arenites. Reidel, Dordrecht, pp. 191-211. Matter, A., Burley, S.D. and Mullis, J., 1985. Quantitative constraints on the timing of reservoir diagenesis: application of combined cathodoluminescence microscopy in fluid inclusion thermometry (Abstr.). Gulf Coast Sec. SEPM Found, 6th Annu. Res. Conf., pp. 19-20. McBride, E.F., 1962. Flysch and associated beds of the Martinsburg Formation (Ordovician), central Appalachians. J. Sediment. Petrol., 32: 39-91. McBride, E.F., 1979. Importance of secondary porosity in sandstone to hydrocarbon exploration (Abstr.). Am. Assoc. Petrol. Geol. Bull., 63: 834. McBride, E.F., 1982. Table of diagenetic sequences of sandstones. Unpublished. McBride, E.F., 1984. Rules of sandstone diagenesis

108 related to reservoir quality (expanded abstr.). Gulf Coast Assoc. Geol. Soc. Trans., 34: 137-139. McBride, E.F., 1987a. Compaction of Norphlet sandstones, Rankin County, Mississippi. Gulf Coast Assoc. Geol. Soc. Trans., 37: 399-404. McBride, E.F., 1987b. Diagenesis of the Maxon Sandstone (Early Cretaceous), Marathon Region, Texas: a diagenetic quartzarenite. J. Sediment. Petrol., 57: 98-107. McBride, E.F., 1988. Silicification of carbonate pebbles in a fluvial conglomerate by groundwater. J. Sediment. Petrol., 58: 862-867. McBride, E.F., 1988. Contrasting diagenetic histories of concretions and host rock, Lion Mountain sandstone (Cambrian), Texas: Geol. Soc. Am. Bull., 100: 1803-1810. McBride, E.F. and Folk, R.L., 1979. Features and origin of Italian Jurassic radiolarites deposited on continental crust. J. Sediment. Petrol., 48: 1069-1102. McBride, E.F. and Thomson, A., 1970. The Caballos Novaculite, Marathon Region, Texas. Geol. Soc. Am. Spec. Pap., 122, 129 pp. McBride, E.F., Weidie, A.E. and Wolleben, J.A., 1975. Deltaic and associated deposits of Difunta Group (Late Cretaceous to Paleocene), Parras and La Popa Basins, northeastern Mexico. In: M.L. Broussard (Editor), Deltas Models for Exploration. Houston Geological Society, pp. 485-552. McBride, E.F., Land, L.S., and Mack, L.E., 1987. Diagenesis of eolian and fluvial feldspathic sandstones, Norphlet Formation (Upper Jurassic) Rankin County, Mississippi and Mobile County, Alabama. Am. Assoc. Petrol. Geol. Bull., 71: 1019-1034. McBride, E.F., Land, L.S., Diggs, T.N. and Mack, L.E., 1988. Petrography, stable isotope geochemistry and diagenesis of Miocene sandstones, Vermilion Block 31, offshore Louisiana. Gulf Coast Assoc. Geol. Soc. Trans. In press. Mellon, G.B., 1964. Discriminatory analysis of calciteand smectite-cemented phases of the Mountain Park Sandstone. J. Geol. 72: 786-809. Melvin, J. and Knight, A.S., 1984. Lithofacies, diagenesis and porosity of the Ivishak Formation, Prudhoe Bay area, Alaska. In: D.A. McDonald and R.C. Surdam (Editors), Clasric Diagenesis. Am. Assoc. Petrol. Geol. Mem., 37: 347-365. Milliken, K.L., 1985. Petrology and Burial Diagenesis of Plio-Pleistocene Sediments, Northern Gulf of Mexico. Ph.D. diss., Univ. of Texas at Austin, 112 PP. Milliken, K.L., Land, L.S. and Loucks, R.G., 1981. History of burial diagenesis determined from isotopic geochemistry, Frio Formation, Brazoria County, Texas. Am. Assoc. Petrol. Geol. Bull., 65: 1397-1413. Millot, G., 1960. Silice, silex, silicifications et croissance des cristaux. Bull. Sere. Carte Geol. Als. Lorr., 13: 129-146. Millot, G., 1970. Geology of Clays. Springer-Verlag, Berlin, 429 pp. Millot, G., Jacques, L. and Wrey, R., 1963. Research in evolution of clay minerals and argillaceous and siliceous neoformation. Clays Clay Miner., 10th Conf., Pergamon Press, MacMillan Co., New York, pp. 399-412. Milnes, A.R. and Thiry, M., 1986. Silicification in Cainozoic sequences in arid northern South Australia (Abstr.). 12th Int. Sediment. Congr., Canberra, Abstracts, p. 213. Molenaar, N., 1986. The interrelation between clay infiltration, quartz cementation, and compaction in Lower Givetian terrestrial sandstones, northern Ardennes, Belgium. J. Sediment. Petrol., 56: 359-369. Moncure, G.K., Lahann, R.W., and Siebert, R.M., 1984. Origin of secondary porosity and cement distribution in a sandstone/shale sequence from the Frio Formation (Oligocene). In: D.A. McDonald and R.C. Surdam (Editors), Clastic Diagenesis. Am. Assoc. Petrol. Geol., 37: 151-161. Morton, J.P., 1985. Rb-Sr evidence for punctuated illite/smectite diagenesis in the Oligocene Frio Formarion, Texas Gulf Coast. Geol. Soc. Am. Bull., 96: 114-122. Nassau, K. and Prescott, B.E., 1977. Smoky, blue, greenish yellow and other irradiarion-related colors in quartz. Mineral. Mag., 41: 301-312. Nanz, R.H., 1954. Genesis of Oligocene Sandstone Reservoir, Seeligson Field, Jim Wells and Kleberg Counties. Am. Assoc. Petrol. Geol. Bull., 38: 96-117. Netto, A.S.T., 1974. Petroleum and Reservoir Potentialities of the Aqua Grande Member (Cretaceous), Reconcavo Basin, Brazil. Master's Thesis, University of Texas, Austin, Texas, 143 pp. Odom, I.E., Willand, T.N. and Lassin, R.J., 1979. Paragenesis of diagenetic minerals in the St. Peter Sandstone (Ordovician), Wisconsin and Illinois. In: P.A. Scholle and P.R. Schluger (Editors), Aspects of Diagenesis. Soc. Econ. Paleontol. Miner. Spec. Publ., 26: 425-443. Okada, H., 1967. Composition and cementation of some Lower Paleozoic grits in Wales. Kyushu Univ. Mem. Fac. Sci., Ser. D, Geol., 18: 261-276. Paraguassu, A.B., 1972. Experimental silicification of sandstone. Geol. Soc. Am. Bull., 83: 2853-2858. Payne, T.G. et al., 1952. The arctic slope of Alaska. U.S. Geol. Sure, Oil and Gas Investig. Map OM 126, Sheet 2. Percival, C.J., 1983. A definition of the term ganister. Geol. Mag., 120: 187-190. Peterson, M.N.A. and von der Borch, C.C., 1965. Chert: modern inorganic deposition in a carbonate-precipitating locality. Science, 149: 1501-1503.

109 Pettijohn, F.J., 1957. Sedimentary Rocks. Harper, New York, N.Y., 719 pp. Pettijohn, F.J., 1975. Sedimentary Rocks. Harper, New York, N.Y., 628 pp. Pettijohn, F.J., Potter, P.E. and Siever, R., 1987. Sand and Sandstone. Springer-Verlag, New York, 553 pp. Pfeiffer, D.S., 1988. Temperature Variations and their Relation to Groundwater Flow, South Texas, Gulf Coast Basin. Master's Thesis, University of Texas, Austin, Texas, 199 pp. Picard, M.D., 1966. Petrography of Red Peak Member, Chugwater Formation (Triassic), west-central Wyoming. J. Sediment. Petrol., 36: 904-926. Pittman, E.D., 1972. Diagenesis of quartz in sandstones as revealed by scanning electron microscopy. J. Sediment. Petrol., 42: 507-519. Pittman, E.D., 1979. Recent advances in sandstone diagenesis. Am. Rev. Earth Planet. Sci., 7: 39-62. Pittman, E.D. and King, G.E., 1986. Petrology and formation damage control. Upper Cretaceous sandstone, offshore Gabon. Clay Miner., 21: 781-790. Pittman, E.D. and Larese, R.E., 1987. Experimental compaction of lithic sands. Soc. Econ. Paleontol. Miner., 6th Annu. Mtg., Abstr., 4: 66. Pittman, E.D. and Lumsden, D.N., 1968. Relationship between chlorite coatings on quartz grains and porosity, Spiro Sand, Oklahoma. J. Sediment. Petrol., 38: 668-671. Porter, E.W. and James, W.C., 1986. Influence of pressure, salinity, temperature and grain size on silica diagenesis in quartzose sandstones. Chem. Geol., 52: 259-369. Putnis, A. and McCormell, J.D.C., 1980. Principles of Mineral Behaviour. Elsevier, New York, N.Y., 257 PP. Prezbindowski, D.R. and Pittman, E.D., 1987. Petrology and geochemistry of sandstones in the Lower Cretaceous submarine fan complex, Deep Sea Drilling Project Hole 603B. In: S.W. van Hinte and J. Wise et al. (Editors), Initial Reports of the Deep Sea Drilling Project, 92: 961-976. Prozorovich, G.E., 1970. Determination of the time of oil and gas accumulation by epigenesis studies. Sedimentology, 15: 41-52. Pye, W.D., 1944. Petrology of Bethel Sandstone of south-central Illinois. Am. Assoc. Petrol. Geol. Bull., 28: 63-122. Pye, K. and Krinsley, D.H., 1985. Formation of secondary porosity in sandstones by quartz framework grain dissolution. Nature, 317: 54-55. Raam, A., 1968. Petrology and diagenesis of Broughton sandstone (Permian), Kiama district, New South Wales. J. Sediment. Petrol., 38: 319-331. Renton, J.J., Heald, M.T., and Cecil, C.B., 1969. Experimental investigation of pressure solution of quartz. J. Sediment. Petrol., 39: 1107-1117. Riezebos, P.A., 1974. Scanning electron microscopical observations on weakly cemented Miocene sands. Geol. Mijnbouw, 53: 109-122. Robinson, G.D., Jr., 1980. Possible quartz synthesis during weathering of quartz-free mafic rock, Jasper County, Georgia. J. Sediment. Petrol., 50: 193-203. Roedder, E., 1979. Fluid inclusion evidence in the environments of sedimentary diagenesis, a review. In: P.A. Scholle and P.R. Schluger (Editors), Aspects of Diagenesis. Soc. Econ. Paleontol. Miner. Spec. Publ., 26: 89-108. Satin, D.D., 1963. Petrography and origin of Peguan sandstones exposed at Chauk, Burma. J. Sediment. Petrol., 33: 333-342. Scherer, M., 1987. Parameters influencing porosity in sandstones: a model for sandstone porosity prediction. Am. Assoc. Petrol. Geol. Bull., 71: 485-491. Schmidt, V., 1976. Secondary porosity in the Parsons Lake Sandstones. Geol. Assoc. Can. Progr. Abstr., 1: 50. Schwab, F.L., 1970. Origin of the Antietam Formation (Late Precambrian-Lower Cambrian), central Virginia. J. Sediment. Petrol., 40: 354-366. Schwab, F.L., 1971. Harpers Formation, central Virginia: a sedimentary model. J. Sediment. Petrol., 41: 139-149. Sears, S.O., 1984. Porcellaneous cement and microporosity in California Miocene turbidites--origin and effect on reservoir properties. J. Sediment. Petrol., 54: 159-169. Sharma, G.D., 1965. Formation of silica cement and its replacement by carbonates. J. Sediment. Petrol., 35: 733-745. Sharp, J.M., Jr., Galloway, W.E., Land, L.S., McBride, E.F. Blanchard, P.E., Bodner, D., Dutton, S.P., Farr, M.R., Gold, P.B., Jackson, T.J., Lundegard, P.D., Milliken, K.L. and Maepherson, G.L., 1989. Diagenetic Processes in Northwest Gulf of Mexico Sediments. Elsevier, Amsterdam. In press. Sibley, D.F. and Blatt, H., 1974. Quantitative evaluation of intergranular pressure solution as a source of silica in the Tuscarora orthoquartzite, central Appalachians. Geol. Soc. Am. Annu. Mtg., 6: 954. Sibley, D.F. and Blatt, H., 1976. Intergranular pressure solution and cementation of the Tuscarora orthoquartzite. J. Sediment. Petrol., 46: 881-896. Siever, R., 1957. Pennsylvanian sandstones of the Eastern Interior Coal Basin. J. Sediment. Petrol., 27: 227-250. Siever, R., 1959. Petrology and geochemistry of silica cementation in some Pennsylvanian sandstones. In: H.A. Ireland (Editor), Silica in Sediments. Soc. Econ. Paleontol. Miner. Spec. Publ., 7: 55-79. Siever, R., 1962. Silica solubility, 0 - 2 0 0 C and the diagenesis of siliceous sediments. J. Geol., 70: 127-151.

110 Siever, R., 1971. Clay diagenesis and sandstone cementation. VIII Int. Sediment. Congr., Progr. Abstr., pp. 93-94. Siever, R., 1979. Plate-tectonics controls on diagenesis. J. Geol., 87: 127-155. Simonson, B.M., 1987. Early silica cementation and subsequent diagenesis in arenites from four Early Proterozoic iron formations of North America. J. Sediment. Petrol., 57: p. 494-511. Sippel, R.F., 1968. Sandstone petrology: evidence from luminescence petrography. J. Sediment. Petrol., 38: 530-554. Smale, D., 1973. Silcretes and associated silica diagenesis in southern Africa and Australia. J. Sediment. Petrol., 43: 1077-1089. Smith, G.W., 1985. Geology of the Deep Tuscaloosa (Upper Cretaceous) Gas Trend in Louisiana. In: B.F. Perkins and G.B. Martin, (Editors), Habitat of Oil and Gas in the Gulf Coast. Gulf Coast Sec. Soc. Econ. Paleontol. Mineral. Found., 4th Annu. Res. Conf., pp. 153-190. Sorby, H.C., 1880. On the structure and origin of noncalcareous stratified rocks. Q. J. Geol. Soc. London, 37: 49-92. Sprunt, E.S., 1981. Causes of quartz cathodoluminescence colors. Scanning Electron Microscopy, Part 1, pp. 525-535. Sprunt, E.S. and Nur, A., 1976. Reduction of porosity by pressure solution: Experimental verification. Geology, 4: 463-466. Stackelberg, P.E., 1986. The Role of Intergranular Pressure Solution In the Diagenesis of the St. Peter Sandstone along a Traverse of the Illinois Basin. Master's Thesis, University of Missouri, Columbia, Missouri, 113 pp. Stalder, P.J., 1975. Cementation of Plio-Quaternary Fluviatile Clastic Deposits in and along the Oman Mountains. Geol. Mijnbouw, 54: 148-156. Stanley, K.O. and Benson, L.V., 1979. Early diagenesis of high plains Tertiary vitric and arkosic sandstone, Wyoming and Nebraska. Soc. Econ. Paleont. Miner. Spec. Publ., 26: 401-423. Stanton, G.D., 1977. Factors Influencing Porosity and Permeability, Wilcox Group (Eocene), Karnes County, Texas. Master's Thesis, Univ. Texas at Austin, Texas, 158 pp. Stephan, H.J., 1970. Diagenesis of the Middle Buntsandstein in South Oldenburg, Lower Saxony, Meyniana, 20: 39-82. Stonecipher, S.A., Winn, R.D., Jr. and Bishop, M.G., 1984. Diagenesis of the Frontier Formation, Moxa Arch: a function of sandstone geometry, texture and composition, and fluid flux. In: D.A. McDonald and R.C. Surdam (Editors), Clastic Diagenesis. Am. Assoc. Petrol. Geol. Mem., 37: 289-316. Suchecki, R.K. and Bloch, S., 1988. Complex quartz overgrowths as revealed by microprobe cathodoluminescence (Abstr.). Am. Assoc. Petrol. Geol. Bull. Sullivan, K.B., 1988. Sandstone and Shale Diagenesis of the Frio Formation (Oligocene), Texas Gulf Coast: A Close Look at Sandstone/Shale Contacts. Master's Thesis, University of Texas, Austin, Texas, 242 pp. Summerfield, M.A., 1983a. Silcrete. In: A.S. Goudie and K. Pye, (Editors), Chemical Sediments and Geomorphology. Academic Press, London, pp. 59-91. Summerfield, M.A., 1983b. Petrography and diagenesis of silcrete from the Kalahari Basin and Cape Coastal Zone, southern Africa. J. Sediment. Petrol., 53: 895-909. Summerfield, M.A., 1986. Silcrete in the sedimentary record: identification, occurrence and palaeoenvironmental significance. 12th Int. Sediment. Congr., Canberra, Abstr., p. 290. Surdam, R.C. and Boles, J.R., 1979. Diagenesis of volcanic sandstones. In: P.A. Scholle and P.R. Schluger (Editors), Aspects of Diagenesis. Soc. Econ. Paleontol. Miner. Spec. Publ., 26: 227-242. Surdam, R.C. and Crossey, L.J., 1987. Integrated diagenetic modeling: a process-oriented approach for elastic systems. Am. Rev. Earth Planet. Sci., 15: 141-170. Surdam, R.C., Boese, S.W. and Crossey, L.J., 1984. The chemistry of secondary porosity. In: D.A. McDonald and R.C. Surdam (Editors), Clastic Diagenesis. Am. Assoc. Petrol. Geol. Mem., 37: 127-149. Swineford, A. and Franks, P.C., 1959. Opal in the Ogallala Formation in Kansas. In: H.A. Ireland (Editor), Silica in Sediments. Soc. Econ. Paleontol. Miner. Spec. Publ., 7: 111-120. Tallman, S.L., 1949. Sandstone types: their abundance and cementing agents. J. Geol., 57: 582-591. Tardy, Y. and Monnin, C., 1982. Recherches sur les mrcanismes du concr&ionnement. Coefficients d'activit6 des ions, solubilit6 des sels et de la silice dans les pores de petite taille. Bull. Mineral., 106: 321-328. Taylor, J.M., 1950. Pore-space reduction in sandstones. Am. Assoc. Petrol. Geol. Bull., 34: 701-716. Thiry, M., Bertrand Ayrault, M., and Grisoni, J-.C., 1988. Ground-water silicification and leaching in sands: example of the Fountainebleau Sand (Oligocene) in the Paris Basin. Geol. Soc. Am. Bull., 100: 1283-1290. Thiry, M. and Millot, G., 1987. Mineralogical forms of silica and their sequence of formation in silcretes. J. Sediment. Petrol., 57: 343-352. Thomas, M.B. and Oliver, T.A., 1979. Depth-porosity relationships in the Viking and Cardium formations of central Alberta. Bull. Can. Petrol. Geol., 27: 209-227.

111 Thomson, A., 1959. Pressure solution and porosity. In: H.A. Ireland (Editor), Silica in Sediments. Soc. Econ. Paleontol. Miner. Spec. Publ., 7: 92-111. Thomson, A., 1978a. Petrology and diagenesis of the Hosston Sandstone reservoirs at Bassfleld, Jefferson Davis County, Mississippi. Gulf Coast Assoc. Geol. Soc. Trans., 28: 651-664. Thomson, A., 1978b. Petrography and diagenesis of Tuscaloosa Sandstone reservoirs at Basefield, Jefferson Davis County, Mississippi (Abstr.). Am. Assoc. Petrol. Geol. Bull., 62: 1769. Tieh, T.T., Berg, R.R., Popp, R.K., Brasher, J.E. and Pike, J.D., 1986. Deposition and diagenesis of upper Miocene arkoses, Yowlumne and Rio Viejo fields, Kern County, California. Am. Assoc. Petrol. Geol. Bull., 70: 953-969. Tillman, R.W. and Almon, W.R., 1979. Diagenesis of Frontier Formation offshore bar sandstones, Spearhead Ranch Field, Wyoming. In: P.A. Scholle and P.R. Schluger (Editors), Aspects of Diagenesis. Soc. Econ. Paleontol. Miner. Spec. Publ., 26: 337-378. Todd, T.W., 1963. Post-depositional history of Tensleep Sandstone (Permsylvanian), Big Horn Basin, Wyoming. Am. Assoc. Petrol. Geol. Bull., 47: 599-616. T6th, J., 1978. Gravity-induced cross-formational flow of formational fluids, Red Earth region, Alberta, Canada. Analysis, patterns, and evolution. Water Resour. Res., 14: 805-843. Towe, K.M., 1962. Clay mineral diagenesis of a possible source of silica cement in sedimentary rocks. J. Sediment. Petrol., 32: 26-28. Travena, A.S. and Clark, R.A., 1986. Diagenesis of sandstone reservoirs of Pattani Basin, Gulf of Thailand. Am. Assoc. Petrol. Geol. Bull., 70: 299-308. Travis, J.W., 1963. Origin of silica cement in quartose rocks: a review. Compass, 41: 5-13. Van Hise, C.R., 1904. A treatise on metamorphism. U.S. Geol. Surv. Monogr., 47, 1286 pp. Von Engelhardt, W., 1967. Interstitial solution and diagenesis in sediments. In: G. Larsen and G.V. Chillingar (Editors), Diagenesis in Sediments. Developments in Sedimentology 8, Elsevier, Amsterdam, pp. 503-522. Waldschmidt, W.A., 1941. Cementing materials in sandstones and their influence on the migration of oil. Am. Assoc. Petrol. Geol. Bull., 25: 1839-1879. Walker, J.R., 1976. Diagenetic origin of continental red beds. In: R. Falke (Editor), The Continental Permian in Central, West, and South Europe. Reidel, Dordrecht, pp. 240-282. Walker, T.R., 1960. Carbonate replacement of detrital crystalline silicate minerals as a source of authigenic silica in sedimentary rocks. Geol. Soc. Am. Bull., 71: 145-152. Walker, T.R., 1967. Formation of red beds in modern and ancient deserts. Geol. Soc. Am. Bull., 78: 353-368. Watts, S.H., 1978. A petrographic study of silcrete from inland Australia. J. Sediment. Petrol., 48: 987-994. Waugh, B., 1970. Formation of quartz overgrowths in the Peurith sandstone (Lower Permian) of northwest England as revealed by scanning electron microscopy. Sedimentology, 14: 309-320. Westcott, W.A., 1983. Diagenesis of Cotton Valley Sandstone (Upper Jurassic), East Texas: Implications for tight gas formation pay recognition. Am. Assoc. Petrol. Geol. Bull., 67: 1002-1013. Weyl, P.K., 1959. Pressure solution and the force of crystallization--A phenomenologlcal theory. J. Geophys. Res., 64: 2001-2005. Whalley, W.B. and Krinsley, D.H., 1974. A scanning electron microscope study of surface textures of quartz grains from glacial environments. Sedimentology, 21: 87-105. Willey, J.D., 1974. The effect of pressure on the solubility of amorphous silicas in seawater at 0 o C. Marine Chem., 2: 239-250. Wilson, J.C. and McBride, E.F., 1988. Compaction and porosity evolution of Pliocene sandstones, Ventura Basin, California, Am. Assoc. Petrol. Geol. Bull., 72: 664-681. Wilson, M.D. and Pittman, E.D., 1977. Authigenic clays in sandstone: recognition and influence on reservoir properties and paleoenvironment analysis. J. Sediment Petrol., 47: 3-31. Wilson, T.V. and Sibley, D.F., 1978. Pressure solution and porosity reduction in a shallow buried quartzarenite. Am. Assoc. Petrol. Geol. Bull., 62: 2329-2334. Wollast, R., 1971. Kinetic aspects of the nucleation and growth of calcite from aqueous solutions. In O.P. Bricker (Editor), Carbonate Cements, John Hopkins Univ. Stud. Geol., 79: 264-273. Wood, J.R. and Hewett, T.A., 1982. Fluid convection and mass transfer in porous sandstone--a theoretical model. Geochim. Cosmochim. Acta, 461707-1713. Wood, J.R. and Hewett, T.A., 1984. Reservoir diagenesis and convective fluid flow. In: D.A. McDonald and R.C. Surdam (Editors), Clastic Diagenesis. Am. Assoc. Petrol. Geol. Mem., 37: 99-110. Wood, J.R. and Surdam, R.C., 1979. Application of convective-diffusion model to diagenetic processes. In: P.A. Scholle and P.R. Schluger (Editors), Aspects of Diagenesis. Soc. Econ. Paleontol. Miner. Spec. Publ., 26: 243-250. Wopfner, H., 1978. Silcretes in northern South Australia and adjacent regions. In: T. Langford-Smith (Editor), Silcrete in Australia. University of New England, Armidale, pp. 93-141. Wright, T.O. and Platt, L.B., 1982. Pressure solution

112 and cleavage in the Martinsburg shale. Am. J. Sci., 282: 122-135. Yeh, H. and Savin, S.M., 1977. Mechanism of burial metamorphism of argillaceous sediments, 3. O-Isotope evidence. Geol. Soc. Am. Bull., 88: 1321-1330. Yurkova, R.M., 1970. Comparison of post-sedimentary alterations of oil-, gas-, and water-bearing rocks. Sedimentology, 15: 53-68. Zinkernagel, U., 1978. Cathodoluminescence of quartz and its application to sandstone petrology. Contrib. Sediment., 8, 69 pp. Zen, E.-An., 1959. Clay mineral-carbonate relations in sedimentary rocks. Am. J. Sci., 257: 29-43. [Received February 25, 1988; accepted after revision August 3, 1988]

You might also like