sedimentary rock

sedimentary rock
Rock formed at or near the Earth's surface by the accumulation and lithification of fragments of preexisting rocks or by precipitation from solution at normal surface temperatures.

Sedimentary rocks can be formed only where sediments are deposited long enough to become compacted and cemented into hard beds or strata. They are the most common rocks exposed on the Earth's surface but are only a minor constituent of the entire crust. Their defining characteristic is that they are formed in layers. Each layer has features that reflect the conditions during deposition, the nature of the source material (and, often, the organisms present), and the means of transport. See also sedimentary facies.

* * *

Introduction

      rock formed at or near the Earth's surface by the accumulation and lithification of sediment (detrital rock) or by the precipitation from solution at normal surface temperatures (chemical rock). Sedimentary rocks are the most common rocks exposed on the Earth's surface but are only a minor constituent of the entire crust, which is dominated by igneous and metamorphic rocks.

      Sedimentary rocks are produced by the weathering of preexisting rocks and the subsequent transportation and deposition of the weathering products. Weathering refers to the various processes of physical disintegration and chemical decomposition that occur when rocks at the Earth's surface are exposed to the atmosphere (mainly in the form of rainfall) and the hydrosphere. These processes produce soil, unconsolidated rock detritus, and components dissolved in groundwater and runoff. erosion is the process by which weathering products are transported away from the weathering site, either as solid material or as dissolved components, eventually to be deposited as sediment. Any unconsolidated deposit of solid weathered material constitutes sediment. It can form as the result of deposition of grains from moving bodies of water or wind, from the melting of glacial ice, and from the downslope slumping (sliding) of rock and soil masses in response to gravity, as well as by precipitation of the dissolved products of weathering under the conditions of low temperature and pressure that prevail at or near the surface of the Earth.

      Sedimentary rocks are the lithified equivalents of sediments. They typically are produced by cementing, compacting, and otherwise solidifying preexisting unconsolidated sediments. Some varieties of sedimentary rock, however, are precipitated directly into their solid sedimentary form and exhibit no intervening existence as sediment. Organic reefs and bedded evaporites are examples of such rocks. Because the processes of physical (mechanical) weathering and chemical weathering are significantly different, they generate markedly distinct products and two fundamentally different kinds of sediment and sedimentary rock: (1) terrigenous clastic sedimentary rocks and (2) allochemical and orthochemical sedimentary rocks.

      Clastic terrigenous sedimentary rocks consist of rock and mineral grains, or clasts, of varying size, ranging from clay-, silt-, and sand- up to pebble-, cobble-, and boulder-size materials. These clasts are transported by gravity, mudflows, running water, glaciers, and wind and eventually are deposited in various settings (e.g., in desert dunes, on alluvial fans, across continental shelves, and in river deltas). Because the agents of transportation commonly sort out discrete particles by clast size, terrigenous clastic sedimentary rocks are further subdivided on the basis of average clast diameter. Coarse pebbles, cobbles, and boulder-size gravels lithify to form conglomerate and breccia; sand becomes sandstone; and silt and clay form siltstone, claystone, mudrock, and shale.

      Chemical sedimentary rocks form by chemical and organic reprecipitation of the dissolved products of chemical weathering that are removed from the weathering site. Allochemical sedimentary rocks, such as many limestones and cherts, consist of solid precipitated nondetrital fragments (allochems) that undergo a brief history of transport and abrasion prior to deposition as nonterrigenous clasts. Examples are calcareous or siliceous shell fragments and oöids, which are concentrically layered spherical grains of calcium carbonate. Orthochemical sedimentary rocks, on the other hand, consist of dissolved constituents that are directly precipitated as solid sedimentary rock and thus do not undergo transportation. Orthochemical sedimentary rocks include some limestones, bedded evaporite deposits of halite, gypsum, and anhydrite, and banded iron formations.

      Sediments and sedimentary rocks are confined to the Earth's crust, which is the thin, light outer solid skin of the Earth ranging in thickness from 40–100 kilometres (25 to 62 miles) in the continental blocks to 4–10 kilometres in the ocean basins. Igneous and metamorphic rocks constitute the bulk of the crust. The total volume of sediment and sedimentary rocks can be either directly measured using exposed rock sequences, drill-hole data, and seismic profiles or indirectly estimated by comparing the chemistry of major sedimentary rock types to the overall chemistry of the crust from which they are weathered. Both methods indicate that the Earth's sediment-sedimentary rock shell forms only about 5 percent by volume of the terrestrial crust, which in turn accounts for less than 1 percent of the Earth's total volume. On the other hand, the area of outcrop and exposure of sediment and sedimentary rock comprises 75 percent of the land surface and well over 90 percent of the ocean basins and continental margins. In other words, 80–90 percent of the surface area of the Earth is mantled with sediment or sedimentary rocks rather than with igneous or metamorphic varieties. The sediment-sedimentary rock shell forms only a thin superficial layer. The mean shell thickness in continental areas is 1.8 kilometres; the sediment shell in the ocean basins is roughly 0.3 kilometre. Rearranging this shell as a globally encircling layer (and depending on the raw estimates incorporated into the model), the shell thickness would be roughly 1–3 kilometres.

      Despite the relatively insignificant volume of the sedimentary rock shell, not only are most rocks exposed at the terrestrial surface of the sedimentary variety, but many of the significant events in Earth history are most accurately dated and documented by analyzing and interpreting the sedimentary rock record instead of the more voluminous igneous and metamorphic rock record. When properly understood and interpreted, sedimentary rocks provide information on ancient geography, termed paleogeography. A map of the distribution of sediments that formed in shallow oceans along alluvial fans bordering rising mountains or in deep, subsiding ocean trenches will indicate past relationships between seas and landmasses. An accurate interpretion of paleogeography and depositional settings allows conclusions to be made about the evolution of mountain systems, continental blocks, and ocean basins, as well as about the origin and evolution of the atmosphere and hydrosphere. Sedimentary rocks contain the fossil record of ancient life-forms that enables the documentation of the evolutionary advancement from simple to complex organisms in the plant and animal kingdoms. Also, the study of the various folds or bends and breaks or faults in the strata of sedimentary rocks permits the structural geology or history of deformation to be ascertained.

      Finally, it is appropriate to underscore the economic importance of sedimentary rocks. For example, they contain essentially the world's entire store of oil and natural gas, coal, phosphates, salt deposits, groundwater, and other natural resources.

      Several subdisciplines of geology deal specifically with the analysis, interpretation, and origin of sediments and sedimentary rocks. Sedimentary petrology is the study of their occurrence, composition, texture, and other overall characteristics, while sedimentology emphasizes the processes by which sediments are transported and deposited. Sedimentary petrography involves the classification and study of sedimentary rocks using the petrographic microscope. stratigraphy covers all aspects of sedimentary rocks, particularly from the perspective of their age and regional relationships as well as the correlation of sedimentary rocks in one region with sedimentary rock sequences elsewhere. (For further information about these fields, see geologic sciences (Earth sciences).)

Classification systems
      In general, geologists have attempted to classify sedimentary rocks on a natural basis, but some schemes have genetic implications (i.e.,knowledge of origin of a particular rock type is assumed), and many classifications reflect the philosophy, training, and experience of those who propound them. No scheme has found universal acceptance, and discussion here will centre on some proposals.

       Terms designating composition and physical characteristicsThe book Rocks and Rock Minerals by Louis V. Pirsson (Pirsson, Louis Valentine) was first published in 1908, and it has enjoyed various revisions. Sedimentary rocks are classified there rather simplistically according to physical characteristics and composition into detrital and nondetrital rocks (see the Table (Terms designating composition and physical characteristics)).

 Numerous other attempts have been made to classify sedimentary rocks. The most significant advance occurred in 1948 with the publication in the Journal of Geology of three definitive articles by the American geologists Francis J. Pettijohn, Robert R. Shrock, and Paul D. Krynine. Their classifications provide the basis for all modern discussion of the subject. The nomenclature associated with several schemes of classifying clastic and nonclastic rocks will be discussed in the following sections, but a rough division of sedimentary rocks based on chemical composition is shown in Figure 1—>.

      For the purposes of the present discussion, three major categories of sedimentary rocks are recognized: (1) terrigenous clastic sedimentary rocks, (2) carbonates (carbonate rock) (limestone and dolomite), and (3) noncarbonate chemical sedimentary rocks. Terrigenous clastic sedimentary rocks are composed of the detrital fragments of preexisting rocks and minerals and are conventionally considered to be equivalent to clastic sedimentary rocks in general. Because most of the clasts are rich in silica, they are also referred to as siliciclastic sedimentary rocks. Siliciclastics are further subdivided on the basis of clast diameter into conglomerate and breccia, sandstone, siltstone, and finer-than-silt-sized mudrock (shale, claystone, and mudstone). The carbonates, limestones and dolomites, consist of the minerals aragonite, calcite, and dolomite. They are chemical sedimentary rocks in the sense that they possess at least in part a crystalline, interlocking mosaic of precipitated carbonate mineral grains. However, because individual grains such as fossil shell fragments exist for some period of time as sedimentary clasts, similar to transported quartz or feldspar clasts, most carbonates bear some textural affinities to the terrigenous clastic sedimentary rocks. The noncarbonate chemical sedimentary rocks include several rock types that are uncommon in the sedimentary rock record but remain important either from an economic point of view or because their deposition requires unusual settings. Specific varieties discussed below include siliceous rocks (cherts), phosphate rocks (phosphorites), evaporites, iron-rich sedimentary rocks (iron formations and ironstones), and organic-rich (carbonaceous) deposits in sedimentary rocks (coal, oil shale, and petroleum).

      Despite the diversity of sedimentary rocks, direct measurement of the relative abundance of the specific types based on the study of exposed sequences suggests that only three varieties account for the bulk of all sedimentary rocks: mudrock, 47 percent; sandstone, 31 percent; and carbonate, 22 percent. Another method, which involves comparing the chemical composition of major sedimentary rock types with the chemistry of the Earth's continental crust, yields somewhat different numbers: mudrock, 79 percent; sandstone, 13 percent; and carbonate, 8 percent. Most sedimentary petrologists concede that the sedimentary rock record preserved and exposed within the continental blocks is selectively biased in favour of shallow-water carbonates and sandstones. Mudrocks are preferentially transported to the ocean basins. Consequently, indirect estimates based on chemical arguments are probably more accurate.

Terrigenous clastic rocks
      A prominent physical feature of terrigenous clastic rocks is texture—that is, the size, shape, and arrangement of the constituent grains. These rocks have a fragmental texture: discrete grains are in tangential contact with one another. Terrigenous clastic sedimentary rocks are further subdivided on the basis of the mean grain diameter that characterizes most fragments, using the generally accepted size limits. Granules, pebbles, cobbles, boulders, and blocks constitute the coarse clastic sediments; sand-size (arenaceous) clasts are considered medium clastic sediments; and fine clastics sediments consists of silt- and clay-size materials.

      The simplest way of classifying coarse clastic sedimentary rocks is to name the rock and include a brief description of its particular characteristics. Conglomerates (conglomerate) and breccias (breccia) differ from one another only in clast angularity. The former consist of abraded, somewhat rounded, coarse clasts, whereas the latter contain angular, coarse clasts. Thus, a pebble conglomerate is a coarse clastic sedimentary rock whose discrete particles are rounded and range from 4 to 64 millimetres (0.2 to 2.5 inches) in diameter. A more precise description reveals the rock types of the mineral fragments that compose the conglomerate—for example, a granite-gneiss pebble conglomerate.

      Sandstones (sandstone) have long intrigued geologists because they are well exposed, are abundant in the geologic record, and provide an enormous amount of information about depositional setting and origin. Many classification schemes have been developed for sandstones, only the most popular of which are reviewed below. Most schemes emphasize the relative abundance of sand-size quartz, feldspar, and rock fragment components, as well as the nature of the material housed between this sand-size “framework” fraction (see below Sandstones (sedimentary rock)).

      Fine clastics are commonly, but rather simplistically, referred to as mudrocks. Mudrocks actually can include any clastic sedimentary rock in which the bulk of the clasts have diameters finer than 1/16 millimetre. Varieties include siltstone (average grain size between 1/16 and 1/256 millimetre) and claystone (discrete particles are mostly finer than 1/256 millimetre). Mud is a mixture of silt- and clay-size material, and mudrock is its indurated product. shale is any fine clastic sedimentary rock that exhibits fissility, which is the ability to break into thin slabs along narrowly spaced planes parallel to the layers of stratification. Despite the great abundance of the fine clastics, disagreement exists as to what classification schemes are most useful for them, and an understanding of their origin is hindered by analytical complexities (see below Mudrocks (sedimentary rock)).

Carbonate rocks: limestones and dolomites
      Limestones and dolostones (dolomites) make up the bulk of the nonterrigenous sedimentary rocks. Limestones are for the most part primary carbonate rocks. They consist of 50 percent or more calcite and aragonite (both CaCO3). Dolomites are mainly produced by the secondary alteration or replacement of limestones; i.e., the mineral dolomite [CaMg(CO3)2] replaces the calcite and aragonite minerals in limestones during diagenesis. A number of different classification schemes have been proposed for carbonates, and the many categories of limestones and dolomites in the geologic record represent a large variety of depositional settings (see below Limestones and dolomites (sedimentary rock)).

Noncarbonate chemical sedimentary rocks
      Noncarbonate chemical sedimentary rocks differ in many respects from carbonate sedimentary rocks and terrigenous clastic sedimentary rocks, and there is no single classification that has been universally accepted. This is a reflection of the great variation in mineral composition, texture, and other properties of these rock types. Such rocks as ironstones and banded iron formations (limonite, goethite, hematite, siderite, and chamosite), phosphorites, evaporites (rock salt, gypsum, and other salts), siliceous rocks (cherts), and organic-rich (carbonaceous) deposits of oil, natural gas, and coal in sedimentary rocks occur in much less abundance than carbonates and siliciclastic sedimentary rocks, although they may form thick and widespread deposits.

 Classification schemes that incorporate all types of noncarbonate chemical sedimentary rocks do not exist because no triangular or tetrahedral scheme can accommodate all of them. Several of the major types are shown in the tetrahedron in Figure 2—>.

Harold J. Bissell Frederick L. Schwab

Properties of sedimentary rocks

Texture
      Texture refers to the physical makeup of rock—namely, the size, shape, and arrangement (packing and orientation) of the discrete grains or particles of a sedimentary rock. Two main natural textural groupings exist for sedimentary rocks: clastic (or fragmental) and nonclastic (essentially crystalline). Noncarbonate chemical sedimentary rocks in large part exhibit crystalline texture, with individual mineral grains forming an interlocking arrangement. Depositional setting is an insignificant factor in both determining crystal size and altering crystalline texture. The size of crystals is controlled to a greater degree by the rate of precipitation, and their texture is modified by postdepositional recrystallization (reflecting the diagenetic (diagenesis) environment). As a result, little attention is paid to crystalline textures other than providing a simple description of it (for example, coarsely crystalline versus finely crystalline). Also, even though carbonate rocks commonly include allochems that behave as clasts, they too are commonly diagenetically altered. Consequently, only cursory efforts are made to texturally characterize limestones and dolomites. Therefore, the following discussion deals in detail only with the textural techniques applied to terrigenous (siliciclastic) sedimentary rocks.

Grain size (grain size scale)
      Particle size is an important textural parameter of clastic rocks because it supplies information on the conditions of transportation, sorting, and deposition of the sediment and provides some clues to the history of events that occurred at the depositional site prior to final induration. Determining the sizes of the discrete particles that constitute a sedimentary rock can be difficult, particularly if the rock is firmly indurated (cemented, compacted, and lithified). Various methods of measuring grain-size distribution have been devised; likewise several different grade-size schemes exist.

      The size of particulate materials that make up sediments and sedimentary rocks are measured by weighing the proportions that accumulate in a series of wire mesh screen sieves, by visually counting grains with a petrographic microscope, or by determining the rate at which particles of varying diameter accumulate in a water-filled glass cylinder (known as a settling tube).

 The millimetre and phi unit grade scales and terminology given in the Table—> are the standard ones used for sediments and sedimentary rocks. In the millimetre scale, each size grade differs from its predecessor by the constant ratio of 1:2; each size class has a specific class name used to refer to the particles included within it. This millimetre, or Udden-Wentworth, scale is a geometric grain-size scale since there is a constant ratio between class limits. Such a scheme is well suited for the description of sediments because it gives equal significance to size ratios, whether they relate to gravel, sand, silt, or clay. The phi scale is a useful, logarithmic-based modification of the Udden-Wentworth scale. Grain-size diameters in millimetres are converted to phi units using the conversion formula: phi (ϕ) = - log2S, where ϕ is size expressed in phi units and S is the grain size in millimetres. Phi values for grains coarser than one millimetre are negative, while those for grains finer than one millimetre are positive.

      After the grain-size distribution for a given sediment or sedimentary rock has been determined by sieving, microscopic analysis, or use of a settling tube, it can be characterized using standard statistical measures in either of two ways: (1) visual inspection of various types of graphs that plot overall percent abundance versus grain-size diameter (e.g., histograms or bar diagrams, size frequency and cumulative size frequency curves, and probability curves that compare the actual grain-size distribution to a normal straight-line Gaussian distribution) or (2) arithmetic calculations made using diameter values in either millimetres or phi units that are read off the graphic plots and inserted into standard formulas. For siliciclastic sedimentary rocks, the following standard statistical measures are conventionally described for grain-size distributions: (1) mode, the most frequently occurring particle size or size class, (2) median, the midpoint size of any grain-size distribution, (3) mean, an estimate of the arithmetic average particle size, (4) sorting or standard deviation, a measure of the range, scatter, or variation in grain size, (5) skewness, the degree of symmetry or asymmetry of the grain-size distribution, which is in turn a function of the coincidence or noncoincidence of mean, median, and mode, and (6) kurtosis (peakedness) of a grain-size distribution, which compares sorting in the central portion of the population with that in the tails.

      Analysis of grain-size distribution is conducted with the disputed assumption that particular transporting agents and depositional settings (e.g., river delta deposits versus shallow marine longshore-bar sands) impose a distinctive textural “fingerprint” on the sediments they produce. Despite continuing efforts, the success of the various graphic and arithmetic approaches in characterizing grain-size distributions is debatable, as is their reliability in pinpointing ancient depositional settings. The grain-size distribution of sediments in many settings commonly appears to be inherited or to exhibit as much variation within a single environment as between different ones.

Particle shape
      Three different but related properties determine particle shape: form, roundness, and surface texture. Particle form is the overall shape of particles, typically defined in terms of the relative lengths of the longest, shortest, and intermediate axes. Particles can be spherical, prismatic, or bladelike. Roundness or angularity is a measure of the smoothness of particles. Surface texture refers to the presence or absence of small, variously shaped markings (pits, polish, scratches) that may occur on grain surfaces.

      Each of these attributes of particle shape is traditionally measured in a standard fashion for the purpose of identifying the transporting agent and the depositional environment. Form is determined either by painstakingly measuring individual particles in three dimensions or by Fourier shape analysis, which uses harmonics analysis and computer digitizing to provide a precise description of particles in two dimensions. Form alone has limited usefulness in inferring depositional setting but more accurately reflects the mineralogy of the grains involved. Roundness is characterized by visually comparing grains to standard silhouette profiles. It is largely the result of abrasion history, which is controlled by the depositional agent and environment. For example, windblown and surf zone sands are well-rounded, while glacial sands and turbidity current deposits are angular. Particle roundness or angularity also reflects mineralogy (soft minerals are abraded more readily than hard minerals), clast size (coarse particles become rounded more rapidly than do fine ones), and transport distance (sands become more abraded and hence rounder as the distance traveled increases). Particle surfaces can be visually examined for pitting, markings, and polish through the use of a microscope or hand lens, or in some cases, a scanning electron microscope (SEM). Certain surface textures have been genetically linked to specific depositional agents; for example, classic V-shaped percussion marks identify quartz grains of the beach and nearshore zones.

      The fabric of a sedimentary rock controls the rock's porosity and permeability and therefore its ability to hold and/or transmit fluids such as oil and water. The orientation, or lack thereof, of the crystals or grains that make up a sedimentary rock constitutes one aspect of fabric. Genetically, there are two principal varieties of oriented fabrics: primary (or depositional) and secondary (or deformational). Primary fabrics are produced while the sediment is accumulating. For example, river currents and some submarine gravity flows generate sediments whose flaky and prismatic constituent particles have long or short axes parallel with one another to produce an oriented fabric. Secondary fabrics result from a rotation of the constituent elements under stress or from the growth of new elements during diagenesis. Fabrics in coarse clastic sedimentary rocks like conglomerates and sandstones can be determined by measuring and plotting dimensional directions, such as the long axes of pebbles or sand grains. In mudrocks, fabrics can be ascertained by studying the platelike arrangement of mica and clay minerals.

      In addition to orientation, a factor known as packing contributes to a rock's fabric. Packing refers to the distribution of grains and intergranular spaces (either empty or filled with cement or fine-grained matrix) in a sedimentary rock. It is controlled by grain size and shape and by the degree of compaction of a sedimentary rock; in turn it determines the rock's bulk density. A description of packing is generally based on the analysis of thin sections of a sedimentary rock using a petrographic microscope. Particular attention is paid to the number of grain-to-grain contacts (packing proximity) and to comparisons between the sum of the lengths of grains to the total length of a traverse across a thin section (packing density).

Mineralogical and geochemical composition
      Minerals that make up sedimentary rocks are of two principal types—namely, detrital and authigenic. Detrital minerals, such as grains of quartz and feldspar, survive weathering and are transported to the depositional site as clasts. Authigenic minerals, like calcite, halite, and gypsum, form in situ within the depositional site in response to geochemical processes. The chemical compounds that constitute them ultimately are generated by chemical weathering and are transported from the weathering site to the point of precipitation primarily in solution. Clay minerals are abundant in sedimentary rocks, particularly mudrocks, and some are detrital. They may have been produced at the weathering site by the partial decomposition of minerals like feldspar. They are transported as clasts, however, and thus can be regarded simply as fine- to very fine-textured detrital particles. Other clay minerals form authigenically at the site of deposition. Some of the important clay minerals are kaolinite, halloysite, montmorillonite, illite, vermiculite, and chlorite.

 The mean chemical composition of the major varieties of sedimentary rocks exhibits wide variation as shown above in Figure 1—>. Significant contrasts in overall composition among sandstones, carbonates, and mudrocks reflect fundamental differences not only in the mechanisms by which detrital minerals of different sizes are transported and deposited but also in the chemical conditions that permit precipitation of various authigenic minerals.

       diagenesis includes all physicochemical, biochemical, and physical processes (short of metamorphism) that modify sediments in the time between their deposition and their analysis. lithification, the process by which sediment is converted into solid sedimentary rock, is one result of diagenesis. Many diagenetic processes such as cementation, recrystallization, and dolomitization are essentially geochemical processes; others like compaction are fundamentally physical processes. All diagenetic changes occur at the low temperatures and pressures characteristic of surface and near-surface environments. These changes can take place almost immediately after sediment formation, or they can occur hundreds or even millions of years later.

Sedimentary structures
      Sedimentary structures are the larger, generally three-dimensional physical features of sedimentary rocks; they are best seen in outcrop or in large hand specimens rather than through a microscope. Sedimentary structures include features like bedding, ripple marks, fossil tracks and trails, and mud cracks. They conventionally are subdivided into categories based on mode of genesis. Structures that are produced at the same time as the sedimentary rock in which they occur are called primary sedimentary structures. Examples include bedding or stratification, graded bedding, and cross-bedding. Sedimentary structures that are produced shortly after deposition and as a result of compaction and dessication are called penecontemporaneous sedimentary structures. Examples include mud cracks and load casts. Still other sedimentary structures like concretions, vein fillings, and stylolites form well after deposition and penecontemporaneous modification; these are known as secondary structures. Finally, others like stromatolites and organic burrows and tracks, though they may in fact be primary, penecontemporaneous, or even secondary, may be grouped as a fourth category—organic sedimentary structures.

      Considerable attention is paid to the sedimentary structures exhibited by any sedimentary rock. Primary sedimentary structures are particularly useful because their abundance and size suggest the probable transporting and depositional agents. Certain varieties of primary sedimentary structures like cross-bedding and ripple marks display orientations that are consistently related to the direction of current movement. Such structures are referred to as directional sedimentary structures because they can be used to infer the ancient paleocurrent pattern or dispersal system by which a sedimentary rock unit was deposited. Other sedimentary structures are stratigraphic “top and bottom” indicators. For example, the progressive upward decrease in clastic grain size diameters, known as graded bedding, would allow a geologist to determine which way is stratigraphically “up”—i.e., toward the younger beds in a dipping sedimentary bed. The suite (repeated sequence) of sedimentary structures in any single stratigraphic unit is another attribute by which that unit may be physically differentiated from others in the region.

Frederick L. Schwab

      Stratification (or bedding) is expressed by rock layers (units) of a general tabular or lenticular form that differ in rock type or other characteristics from the material with which they are interstratified (sometimes stated as interbedded, or interlayered). These beds, or strata, are of varying thickness and areal extent. The term stratum identifies a single bed, or unit, normally greater than one centimetre in thickness and visibly separable from superjacent (overlying) and subjacent (underlying) beds. “Strata” refers to two or more beds, and the term lamina is sometimes applied to a unit less than one centimetre in thickness. Thus, lamination consists of thin units in bedded, or layered, sequence in a natural rock succession, whereas stratification consists of bedded layers, or strata, in a geologic sequence of interleaved sedimentary rocks.

      For most stratified sedimentary rocks, the arrangement of layers is one of unequal thickness, ranging from very thin laminae to discrete beds that measure a few to many metres in thickness. The terms thick and thin as applied to bedding, or stratification, are relative, reflecting the training of a particular geologist as well as experience with a specific stratigraphic section or sections.

Bedding types and bedding-plane features
Types
      It is common to discover a rhythmic pattern in a pile of stratified sedimentary rocks represented by a repetitive sequence of rock types. In most instances of such cyclic sedimentation, the bedding, or stratification, is horizontal or essentially so; that is, the transporting, sorting, and depositing agents of wind, running water, and lake and ocean currents and waves accumulated the laminae and strata in a flat-lying or horizontal arrangement. They are termed well-bedded, a type of primary stratification.

      Primary stratification in sediments and sedimentary rocks can be cross-bedded (cross-stratified), graded, and imbricate and can also display climbing laminae, ripples, and beds.

 Graded bedding simply identifies strata that grade upward from coarse-textured clastic sediment at their base to finer-textured materials at the top (Figure 3—>). The stratification may be sharply marked so that one layer is set off visibly from those above and beneath it. More commonly, however, the layers are blended. This variety of bedding results from a check in the velocity of the transporting agent, and thus coarse-textured sediment (gravel, for example) is deposited first, followed upward by pebbles, granules, sand, silt, and clay. It is commonly associated with submarine density currents.

 Imbricate bedding is a shingle structure in a deposit of flattened or disk-shaped pebbles or cobbles (Figure 3—>). That is to say, elongated and commonly flattened pebbles and cobbles in gravelly sediment are deposited so that they overlap one another like roofing shingles. Imbricate bedding forms where high-velocity currents move over a streambed or where strong currents and waves break over a gradually sloping beach, thereby forming beach shingle.

      Growth structures in sedimentary rocks are in situ features that accumulate largely as the result of organic buildups within otherwise horizontal or nearly flat-lying strata. Reefs (coral reef) and stromatolites (stromatolite) are two common varieties of such growth structures.

Bedding-plane features
      Upper surfaces of beds commonly display primary sedimentary features that are classified as bedding-plane structures. A three-dimensional view may be obtained if some of these can be seen from the side as well as from the top of a pile of strata. They include such features as ripples (ripple marks), climbing ripples, rills, pits, mud cracks, trails and tracks, salt and ice casts and molds, and others. Bedding-plane markings and irregularities can be allocated to one of three classes: (1) those on the base of a bed (load and current structures and organic markings), (2) those within a bed (parting lineation), and (3) those on top of a bed (ripple marks, pits, impressions, mud cracks, tracks and trails of organisms, and others).

Deformation (deformation and flow) structures
      In addition to sedimentary structures that are normally associated with bedding planes, there are other such structures that result from deformation during or shortly after sedimentation but before induration of the sediment into rock. These are nontectonic features—i.e., they are not bends and folds brought about by metamorphism or other such causes. Deformation structures can be grouped into several classes, as follows: (1) founder and load structures, (2) convoluted structures, (3) slump structures, (4) injection structures, such as sandstone dikes or sills, and (5) organic structures.

      Structures found on the bottom of a bed are called sole markings, because they formed on the “sole” of the bed. Sole marks are commonly formed on sandstone and limestone beds that rest upon shale beds. They are termed casts, because they are fillings of depressions that formed on the surface of the underlying mud. They originate (1) by unequal loading upon the soft and plastic wet mud, (2) by the action of currents across the upper mud surface, or (3) by the activities of organisms on this surface. Load casts form as the result of downsinking of sandstone or limestone into the mud beneath. Current marks can form by the action of water currents on upper surfaces of the beds or by “tools” (such as wood and fossils) that are transported by currents over soft sediment.

Sedimentary environments
      The sedimentary environment is the specific depositional setting of a particular sedimentary rock and is unique in terms of physical, chemical, and biological characteristics. The physical features of a sedimentary environment include water depth and the velocity and persistence of currents. Chemical characteristics of an environment include the salinity (proportion of dissolved salts), acidity or basicity (pH), oxidation potential (Eh), pressure, and temperature. The biological characteristics are mainly the assemblage of fauna and flora that populate the setting. These conditions, combined with the nature of the transporting agent and the source area, largely determine the properties of the sediments deposited within the environment. A number of ways of classifying depositional environments exist, but most modern schemes employ a geomorphologic approach. That is to say, an environment is defined in terms of a distinct geomorphic unit or landform, modern examples of which are readily visible for comparative purposes—e.g., a river delta, an alluvial fan, a submarine fan, or the abyssal floor of an ocean basin.

      Individual environments are further grouped into (1) marine environments, which include the nearshore, shallow littoral zone and the offshore, deep littoral zone, as well as deepwater realms, (2) mixed marine and nonmarine settings such as the beach and supratidal zones, and (3) nonmarine settings like lacustrine and various alluvial settings. Each environment is associated with a set of criteria that constitutes its distinguishing features.

Harold J. Bissell Frederick L. Schwab

Sedimentary rock types

Conglomerates (conglomerate) and breccias
      Conglomerates and breccias (breccia) are sedimentary rocks composed of coarse fragments of preexisting rocks held together either by cement or by a finer-grained clastic matrix. Both contain significant amounts (at least 10 percent) of coarser-than-sand-size clasts. Breccias are consolidated rubble; their clasts are angular or subangular. Conglomerates are consolidated gravel whose clasts are subrounded to rounded. Sometimes the term rudite (or rudaceous) is used to collectively refer to both breccias and conglomerates.

Classification schemes
      A number of classification schemes have been proposed to further subdivide conglomerates and breccias. One scheme is purely descriptive, partitioning these coarse clastic sedimentary rocks on the basis of grain size (e.g., boulder breccia versus cobble conglomerate) or composition or both (chert pebble breccia versus limestone cobble conglomerate). Yet another scheme differentiates individual conglomerates and breccias according to depositional agency and environmental setting (alluvial fan conglomerate as opposed to beach conglomerate). The best classification systems incorporate objective physical characteristics of both composition and texture as well as mode of genesis. Conglomerates and breccias belong to four genetic categories: (1) epiclastic, produced by the physical disintegration (weathering) of preexisting rocks, (2) pyroclastic, produced by the explosive activity of volcanoes, (3) cataclastic, formed by local earth movements (fault breccias) or solution phenomena (collapse breccias), and (4) meteoritic, produced by the impact of extraterrestrial bodies on the Earth's surface. In a strict sense, epiclastic conglomerates and breccias are the only true sedimentary rocks, because they alone are produced by weathering.

Ernst ten Haaf Frederick L. Schwab

Epiclastic conglomerates and breccias
      There are two principal types of epiclastic conglomerates and breccias: intraformational, derived penecontemporaneously by eroding, transporting, and depositing material from within the depositional basin itself; and interformational, derived from source rocks that lie outside the area in which the deposit occurs. Epiclastic conglomerates and breccias together probably make up no more than 1 or 2 percent of the conventional sedimentary rock record.

      Intraformational conglomerates and breccias are widespread in the geologic record but are volumetrically unimportant. They occur as laterally continuous bands or horizons within sequences of shallow-water marine or nonmarine deposits. Their origin is commonly related to the existence of brief episodes of strong bottom-hugging currents capable of ripping up recently deposited, unconsolidated sediment. For example, shallow marine limestone deposits commonly have thin bands of boulder-, cobble-, and pebble-size carbonate clasts (edgewise conglomerate or breccia beds) that are generated when storm waves erode and redeposit carbonate mud layers. Likewise, high-velocity river currents that accompany torrential rains give rise to shale pebble conglomerates and breccias within sequences of floodplain alluvium. Other intraformational conglomerates and breccias mix shallow- and deep-water sedimentary rock clasts encased in a finer-grained matrix of deeper-water material. Such deposits accumulate as depositional aprons that flank the scarps (steep slopes) bounding shallow-water platformal areas such as the modern Great Bahama and Little Bahama banks off Florida in the southeastern United States.

      Interformational conglomerates and breccias, the coarse clastic sedimentary rocks derived by the weathering of preexisting rocks outside the depositional basin, are most important from the points of view of both volume and geologic significance. They can be subdivided into two specific categories: (1) clast-supported conglomerates (and breccias) and (2) matrix-supported conglomerates.

Clast-supported conglomerates
      These rocks contain less than 15 percent matrix—i.e., material composed of clasts finer than granule size (2-millimetre diameter or less). They typically exhibit an intact fabric that has a clast-supported framework such that the individual granules, pebbles, cobbles, and boulders touch each other. The space between framework components either is empty or is filled with chemical cement, finer-than-granule clasts, or both. Clast-supported conglomerates may be composed of large clasts of a single rock or mineral type (oligomictic orthoconglomerates), or they may contain a variety of rock and mineral clast types (petromictic orthoconglomerates). If the clasts are all of quartz, then it is called a quartzose conglomerate.

      Clast-supported conglomerates (and orthobreccias) are deposited by highly turbulent water. For example, beach deposits commonly contain lenses and bands of oligomictic orthoconglomerate, composed mainly (95 percent or more) of stable, resistant, coarse clasts of vein quartz, quartzite, quartz sandstone, and chert. Such deposits are typically generated in the upper reaches of winter storm beaches where strong surf can sift, winnow, and abrade coarse pebbles and boulders. The most indestructible components are thereby consolidated as conglomeratic lenses that are interfingered with finer-grained, quartz-rich beach sandstones. Petromictic conglomerates and breccias, on the other hand, reflect the existence of high-relief (mountainous) source areas. Topographically high source areas signify tectonic mobility in the form of active folding or faulting or both. The existence of petromictic conglomerates and breccias in the geologic record is therefore significant: their presence and age not only pinpoint the timing and location of mountain building accompanied by sharp uplift and the possibility of regionally significant fault scarps but also can be used to infer the past distribution of physiographic features such as mountain fronts, continental block margins, continental shelf-continental slope boundaries, and the distribution of oceanic trenches and volcanic island arcs. Deep marine conglomerates may be called resedimented conglomerates. These were retransported from seashore areas by turbidity currents.

      Bedding (layering and stratification) in clast-supported conglomerates, if apparent at all, is typically thick and lenticular. Graded bedding, in which size decreases from bottom to top, is common: because agitated waters rarely subside at once, declining transport power causes a gradual upward decrease in maximum clast size. Relative to the bedding, the pebbles in sandy conglomerates tend to lie flat, with their smallest dimension positioned vertically and the greatest aligned roughly parallel to the current. In closely packed orthoconglomerates, however, there is often a distinct imbrication; i.e., flat pebbles overlap in the same direction like roof shingles. Imbrication is upstream on riverbeds and seaward on beaches.

      Clast-supported conglomerates are quite important economically because they hold enormous water reserves that are easily released through wells. This feature is attributable to their high porosity and permeability. Porosity is the volume percentage of “void” (actually fluid- or air-filled) space in a rock, whereas permeability is defined as the rate of flow of water at a given pressure gradient through a unit volume. The interconnectedness of voids in conglomerates contributes to their permeability. Also, because the chief resistance to flow is generally due to friction and capillary effects, the overall coarse grain size makes conglomerates even more permeable. The high degree of porosity and permeability causes conglomerates to generate excellent surface drainage, so they are to be avoided as dam and reservoir sites.

Matrix-supported conglomerates
      Matrix-supported conglomerates, also called diamictites, exhibit a disrupted, matrix-supported fabric; they contain 15 percent or more (sometimes as much as 80 percent) sand-size and finer clastic matrix. The coarse detrital clasts “float” in a finer-grained detrital matrix. They actually are mudrocks in which there is a sprinkling of granules, pebbles, cobbles, or boulders, or some combination of them. Accordingly, they are sometimes referred to as conglomerate mudstones or pebbly mudstones.

      Although matrix-supported conglomerates originate in a variety of ways, they are not deposited by normal currents of moving water. Some are produced by submarine landslides, massive slumping, or dense, sediment-laden, gravity-driven turbidity flows. Matrix-supported conglomerates that can be definitively related to such mechanisms are called tilloids. Tilloids commonly make up olistostromes, which are large masses of coarse blocks chaotically mixed within a muddy matrix. The terms till (when unconsolidated) and tillite (when lithified) are used for diamictites that appear to have been directly deposited by moving sheets of glacial ice. Tillites typically consist of poorly sorted angular and subangular, polished and striated blocks of rock floating in an unstratified clay matrix. The clasts may exhibit a weak but distinct alignment of their long axes approximately parallel to the direction of ice flow. Tillites are notoriously heterogeneous in composition: clasts appear to be randomly mixed together without respect to size or compositional stability. These clasts are derived mainly from the underlying bedrock. Extremely coarse, far-traveled blocks and boulders are called erratics.

      Other rarer diamictites, known as laminated pebbly (or cobbly or bouldery) mudstones, consist of delicately laminated mudrocks in which scattered coarser clasts occur. Laminations within the muddy component are broken and bent. They are located beneath and adjacent to the larger clasts but gently overlap or arch over them, suggesting that the coarse clasts are dropstones (i.e., ice-rafted blocks released as floating masses of ice melt).

Sandstones (sandstone)
      Sandstones are siliciclastic sedimentary rocks that consist mainly of sand-size grains (clast diameters from 2 to 1/16 millimetre) either bonded together by interstitial chemical cement or lithified into a cohesive rock by the compaction of the sand-size framework component together with any interstitial primary (detrital) and secondary (authigenic) finer-grained matrix component. They grade, on the one hand, into the coarser-grained siliciclastic conglomerates and breccias described above, and, on the other hand, into siltstones and the various finer-grained mudrocks described below. Like their coarser analogues—namely, conglomerates and breccias—sand-size (also called arenaceous) sedimentary rocks are not exclusively generated by the physical disintegration of preexisting rocks. Varieties of limestone that contain abundant sand-size allochems like oöids and fossil fragments are, in at least a textural sense, types of sandstones, although they are not terrigenous siliciclastic rocks. Such rocks, called micrites (micrite) when lithified or carbonate sands when unconsolidated, are more properly discussed as limestones. Also, pyroclastic sandstones or tuffs formed by lithifying explosively produced volcanic ash deposits can be excluded from this discussion because their origin is unrelated to weathering.

      Sandstones are significant for a variety of reasons. Volumetrically they constitute between 10 and 20 percent of the Earth's sedimentary rock record. They are resistant to erosion and therefore greatly influence the landscape. When they are folded, they create the backbone of mountain ranges like the Appalachians of eastern North America, the Carpathians of east-central Europe, the Pennines of northern England, and the Apennine Range of Italy; when flat-lying, they form broad plains and plateaus like the Colorado and Allegheny plateaus. Sandstones are economically important as major reservoirs for both petroleum and water, as building materials, and as valuable sources of metallic ores. Most significantly, they are the single most useful sedimentary rock type for deciphering Earth history. Sandstone mineralogy is the best indicator of sedimentary provenance: the nature of a sedimentary rock source area, its composition, relief, and location. Sandstone textures and sedimentary structures also are reliable indexes of the transportational agents and depositional setting.

Sandstone components and colour
      There are three basic components of sandstones: (1) detrital grains, mainly transported, sand-size minerals such as quartz and feldspar, (2) a detrital matrix of clay or mud, which is absent in “clean” sandstones, and (3) a cement that is chemically precipitated in crystalline form from solution and that serves to fill up original pore spaces.

      The colour of a sandstone depends on its detrital grains and bonding material. An abundance of potassium feldspar often gives a pink colour; this is true of many feldspathic arenites, which are feldspar-rich sandstones. Fine-grained, dark-coloured rock fragments, such as pieces of slate, chert, or andesite, however, give a salt-and-pepper appearance to a sandstone. Iron oxide cement imparts tones of yellow, orange, brown, or red, whereas calcite cement imparts a gray colour. A sandstone consisting almost wholly of quartz grains cemented by quartz may be glassy and white. A chloritic clay matrix results in a greenish black colour and extreme hardness; such rocks are wackes.

Formation of sandstones today
      Sandstones occur in strata of all geologic ages. Much scientific understanding of the depositional environment of ancient sandstones comes from detailed study of sand bodies forming at the present time. One of the clues to origin is the overall shape of the entire sand deposit. Inland desert sands today cover vast areas as a uniform blanket; some ancient sandstones in beds a few hundred metres thick but 1,600 kilometres or more in lateral extent, such as the Nubian Sandstone of North Africa, of Mesozoic age (about 245 to 66.4 million years old), also may have formed as blankets of desert sand. Deposits from alluvial fans (alluvial fan) form thick, fault-bounded prisms. River sands today form shoestring-shaped bodies, tens of metres thick, a few hundred metres wide, up to 60 kilometres or more long, and usually oriented perpendicularly to the shoreline. In meandering back and forth, a river may construct a wide swath of sand deposits, mostly accumulating on meander-point bars. Beaches, coastal dunes, and barrier bars also form “shoestring” sands, but these are parallel to the shore. Deltaic (delta) sands show a fanlike pattern of radial, thick, finger-shaped sand bodies interbedded with muddy sediments. Submarine sand bodies are diverse, reflecting the complexities of underwater topography and currents. They may form great ribbons parallel with the current; huge submarine “dunes” or “sand waves” aligned perpendicularly to the current; or irregular shoals, bars, and sheets. Some sands are deposited in deep water by the action of density currents, which flow down submarine slopes by reason of their high sediment concentrations and, hence, are called turbidity currents. These characteristically form thin beds interbedded with shales; sandstone beds often are graded from coarse grains at the base to fine grains at the top of the bed and commonly have a clay matrix.

      One of the most fruitful methods of deciphering the environment of deposition and direction of transport of ancient sandstones is detailed field study of the sedimentary structures.

      Bedding in sandstones, expressed by layers of clays, micas, heavy minerals, pebbles, or fossils, may be tens of feet thick, but it can range downward to paper-thin laminations. Flagstone breaks in smooth, even layers a few centimetres thick and is used in paving. Thin, nearly horizontal lamination is characteristic of many ancient beach sandstones. Bedding surfaces of sandstones may be marked by ripples (almost always of subaqueous origin), by tracks and trails of organisms, and by elongated grains that are oriented by current flow (fossils, plant fragments, or even elongated sand grains). Sand-grain orientation tends to parallel direction of the current; river-channel trends in fluvial sediments, wave-backwash direction in beach sands, and wind direction in eolian sediments are examples of such orientation.

      A great variety of markings, such as flutes and scour and fill grooves, can be found on the undersides of some sandstone beds. These markings are caused by swift currents during deposition; they are particularly abundant in sandstones deposited by turbidity currents.

      Within the major beds, cross-bedding is common. This structure is developed by the migration of small ripples, sand waves, tidal-channel large-scale ripples, or dunes and consists of sets of beds that are inclined to the main horizontal bedding planes. Almost all sedimentary environments produce characteristic types of cross-beds; as one example, the lee faces of sand dunes (side not facing the wind) may bear cross-beds as much as 33 metres (108 feet) high and dipping 35°.

      Some sandstones contain series of graded beds. The grains at the base of a graded bed are coarse and gradually become finer upward, at which point there is a sharp change to the coarse basal layer of the overlying bed. Among the many mechanisms that can cause these changes in grain size are turbidity currents, but in general they can be caused by any cyclically repeated waning current.

      After the sand is deposited, it may slide downslope or subside into soft underlying clays. This shifting gives rise to contorted or slumped bedding on a scale of centimetres to tens of metres. Generally these are characteristic of unstable areas of rapid deposition.

      Local cementation may result in concretions of calcite, pyrite, barite, and other minerals. These can range from sand crystals or barite roses to spheroidal or discoidal concretions tens of metres across.

      The fossil content also is a useful guide to the depositional environment of sandstones. Desert sandstones usually lack fossils. River-channel and deltaic sandstones may contain fossil wood, plant fragments, fossil footprints, or vertebrate remains. Beach and shallow marine sands contain mollusks, arthropods, crinoids, and other marine creatures, though marine sandstones are much less fossiliferous than marine limestones. Deepwater sands are frequently devoid of skeletal fossils, although tracks and trails may be common. The fossils are not actually structures, of course, but the living organisms were able to produce them. Burrowing by organisms, for example, may cause small-scale structures, such as eyes and pods or tubules of sand.

      The texture of a sandstone is the sum of such attributes as the clay matrix, the size and sorting of the detrital grains, and the roundness of these particles. To evaluate this property, a scale of textural maturity that involved four textural stages was devised in 1951. These stages are described as follows.

      Immature sandstones contain a clay matrix, and the sand-size grains are usually angular and poorly sorted. This means that a wide range of sand sizes is present. Such sandstones are characteristic of environments in which sediment is dumped and is not thereafter worked upon by waves or currents. These environments include stagnant areas of sluggish currents such as lagoons or bay bottoms or undisturbed seafloor below the zone of wave or current action. Immature sands also form where sediments are rapidly deposited in subaerial environments, such as river floodplains, swamps, alluvial fans, or glacial margins. Submature sandstones are created by the removal of the clay matrix by current action. The sand grains are, however, still poorly sorted in these rocks. Submature sandstones are common as river-channel sands, tidal-channel sands, and shallow submarine sands swept by unidirectional currents. Mature sandstones are clay-free, and the sand grains are subangular, but they are well sorted—that is, of nearly uniform particle size. Typically, these sandstones form in environments of current reversal and continual washing, such as beaches. Supermature sandstones are those that are clay-free and well sorted and, in addition, in which the grains are well rounded. These sandstones probably formed primarily as desert dunes, where intense eolian abrasion over a very long period of time may wear sand grains to nearly spherical shapes.

      The methodology used for detailed study of siliciclastic sedimentary rock textures, particularly grain-size distribution and grain shape (angularity and sphericity) has been described above. The information that results from textural analyses is especially useful in identifying sandstone depositional environments. Dune sands in all parts of the world, for example, tend to be fine-sand-size (clast diameters from 1/4 to 1/8 millimetre) because sand of that dimension is most easily moved by winds. Desert (eolian) sandstones also tend to be bimodal or polymodal—i.e., having two (or more) abundant grain-size classes separated by intervening, less prevalent size classes. Dune and beach sands exhibit the best sorting; river and shallow marine sands are less well sorted. River-floodplain, deltaic, and turbidity-current sand deposits show much poorer sorting. Skewness (the symmetry or asymmetry of a grain-size distribution) also varies as a function of depositional setting. Beach sands are commonly negatively skewed (they have a tail of more poorly sorted coarse grains), whereas dune and river sands tend to be positively skewed (a tail of more poorly sorted fine grains).

      Careful analysis of grain roundness and grain shape also can aid in distinguishing the high-abrasion environments of beach and especially dune sands from those of fluvial or marine sands. Rounding takes place much more rapidly in sands subjected to wind action than in water-laid sands. In general, coarser sand grains are better rounded than finer grains because the coarser ones hit bottom more frequently and also hit with greater impact during transport. Sand grains may also have polished, frosted, pitted, or otherwise characteristic surfaces. These depend on the grain size, the agent of transport, and the amount of chemical attack. For example, polish can occur on medium-grained beach sands and fine-grained desert sands and can also be produced chemically by weathering processes.

Classification of sandstones
      There are many different systems of classifying sandstones, but the most commonly used schemes incorporate both texture (the presence and amount of either interstitial matrix—i.e., clasts with diameters finer than 0.03 millimetre—or chemical cement) and mineralogy (the relative amount of quartz and the relative abundance of rock fragments to feldspar grains). The system presented here (Figure 4) is that of the American petrologist Robert H. Dott (1964), which is based on the concepts of P.D. Krynine and F.J. Pettijohn. Another popular classification is that of R.L. Folk (1974). Although these classifications were not intended to have tectonic significance, the relative proportions of quartz, feldspar, and fragments are good indicators of the tectonic regime. It is possible to discriminate between stable cratons (rich in quartz and feldspar), orogens (rich in quartz and fragments), and magmatic arcs (rich in feldspar and fragments).

      Sandstones are first subdivided into two major textural groups, arenites (arenite) and wackes (wacke). Arenites (the front triangular panel of Figure 4) consist of a sand-size framework component surrounded by pore spaces that are either empty (in the case of arenite sands) or filled with crystalline chemical cement (in the case of arenites). Wackes (the second triangular panel of Figure 4) consist of a sand-size framework component floating in a finer-grained pasty matrix of grains finer than 0.03 millimetre whose overall abundance exceeds 15 percent by volume. A third triangular panel in the background shows the natural transition from sandstones to mudrocks as the percentage of sand-size framework clasts decreases.

      Further subdivision of both arenites and wackes into three specific sandstone families is based on the relative proportions of three major framework grain types: quartz (Q), feldspar (F), and rock fragments (R for rock fragment, or L for lithic fragment). For example, quartz arenites are rocks whose sand grains consist of at least 95 percent quartz. If the sand grains consist of more than 25 percent feldspar (and feldspar grains are in excess of rock fragments), the rock is termed arkosic arenite or “ arkose,” although such sandstones are also somewhat loosely referred to as feldspathic sandstones. In subarkosic arenite (or subarkose), feldspar sand grains likewise exceed rock fragments but range in abundance from 5 to 15 percent. Lithic arenites (lithic arenite) have rock fragments that exceed feldspar grains; the abundance of rock fragments is greater than 25 percent. Sublithic arenites likewise contain more rock fragments than feldspar, but the amount of rock fragments is lower, ranging from 5 to 25 percent. Lithic arenites can be further subdivided according to the nature of the rock fragments, as shown in the smaller triangle of Figure 4. This classification scheme also recognizes three major types of wackes or graywackes that are roughly analogous with the three major arenite groups: quartz wacke, feldspathic wacke (with the subvariety arkosic wacke), and lithic wacke. The three major arenite sandstone families are separately described below, but the varieties of wacke can be conveniently considered together as a single group.

Quartz arenites
      Quartz arenites are usually white, but they may be any other colour; cementation by hematite, for example, makes them red. They are usually well sorted and well rounded (supermature) and often represent ancient dune, beach, or shallow marine deposits. Characteristically, they are ripple-marked or cross-bedded and occur as widespread thin blanket sands. On chemical analysis, some are found to contain more than 99 percent SiO2 (quartz). Most commonly they are cemented with quartz, but calcite and iron oxide frequently serve as cements as well.

      This type of sandstone is widespread in stable areas of continents surrounding the craton, such as central North America (St. Peter Sandstone of Ordovician age [about 505 to 438 million years old]), central Australia, or the Russian Platform, and are particularly common in Paleozoic strata (that formed from 570 to 245 million years ago). Quartz arenites have formed in the past when large areas of subcontinental dimensions were tectonically stable (not subject to uplift or deformation) and of low relief, so that extensive weathering could take place, accompanied by prolonged abrasion and sorting. This process eliminated all the unstable or readily decomposed minerals such as feldspar or rock fragments and concentrated pure quartz together with trace amounts of zircon, tourmaline, and various other resistant heavy minerals.

      Quartz arenites have also accumulated to thicknesses of hundreds and even thousands of metres on the continental shelf areas produced as passive continental margins develop during the early stages of continental rifting and the opening of an ocean basin. These thick, continental margin deposits form only if source areas are sufficiently stable to permit beach abrasion and intense chemical weathering capable of destroying rock fragments and feldspars. Subsequent ocean basin closure and continental collision deforms the continental shelf and rise assemblages, incorporating clean quartz arenite units into the resulting folded and faulted mountain system, typically as major ridges. Examples include the Cambrian Chilhowee Group and Silurian Tuscarora Sandstone and Clinch Sandstone formations in the Appalachian Mountains of eastern North America and the Flathead Sandstone and Tapeats Sandstone of the Rocky Mountains in the western part of the continent.

Arkosic sandstones
      Arkosic sandstones are of two types. The most common of these is a mixture of quartz, potash feldspar, and granitic rock fragments. Chemically, these rocks are 60–70 percent silica (or silicon dioxide) and 10–15 percent aluminum oxide (Al2O3), with significant amounts of potassium (K), sodium (Na), and other elements. This type of arkosic sandstone, or arkose, can form wherever block faulting of granitic rocks occurs, given rates of uplift, erosion, and deposition that are so great that chemical weathering is outweighed and feldspar can survive in a relatively unaltered state. These rocks are usually reddish, generally immature, very poorly sorted, and frequently interbedded with arkose conglomerate; alluvial fans or fluvial aprons are the main depositional environments. The Triassic Newark Group of Connecticut is a classic example of this type of arkosic sandstone.

      Arkoses also form under desert (or rarely Arctic) conditions in which the rate of chemical decomposition of the parent granite or gneiss is very slow. These arkoses are generally well sorted and rounded (supermature) and show other desert features, such as eolian cross-beds, associated gypsum, and other evaporitic minerals. The Precambrian Torridonian Arkose of Great Britain is thought to be of desert origin. Basal sands deposited on a granitic-gneissic craton also are usually arkosic. Subarkose sandstones (e.g., Millstone Grit from the Carboniferous of England) have a feldspar content that is diminished by more extensive weathering or abrasion or by dilution from nonigneous source rocks.

Lithic arenites (lithic arenite)
      Lithic arenites occur in several subvarieties, but they are normally gray or of salt-and-pepper appearance because of the inclusion of dark-coloured rock fragments. Most commonly, fragments of metamorphic rocks such as slate, phyllite, or schist predominate, producing phyllarenite. If volcanic rock fragments such as andesite and basalt are most abundant, the rock is termed a volcanic arenite. If chert and carbonate rock fragments are predominant, the name chert or calclithite is applied.

      Lithic arenites are usually rich in mica and texturally immature; the silicon dioxide content is 60–70 percent; aluminum oxide is 15 percent; and potassium, sodium, iron (Fe), calcium (Ca), and magnesium (Mg) are present in lesser amounts. Lithic arenites are very common in the geologic record, are widespread geographically, and are of all ages. They generally were formed as the result of rapid uplift, intense erosion, and high rates of deposition. Many of the classic postorogenic clastic wedge systems found in the major mountain systems of the world contain abundant lithic arenites. In the Appalachians, these include the Ordovician Juniata Formation of the Taconic clastic wedge, the Devonian Catskill Formation of the Acadian clastic wedge, and the Pocono and Mauch Chunk formations of the Alleghenian clastic wedge. Most lithic arenites are deposited as fluvial apron, deltaic, coastal plain, and shallow marine sandstones, interbedded with great thicknesses of shale and frequently with beds of coal or limestone. If they are deposited in an oxidizing environment such as a well-drained river system, they are reddish (e.g., the Catskill Formation of the northeastern United States and the Devonian Old Red Sandstone of England).

Robert Louis Folk Frederick L. Schwab

Wackes (wacke)
      Wacke, or graywacke (wacke), is the name applied to generally dark-coloured, very strongly bonded sandstones that consist of a heterogeneous mixture of rock fragments, feldspar, and quartz of sand size, together with appreciable amounts of mud matrix. Almost all wackes originated in the sea, and many were deposited in deep water by turbidity currents.

      Wackes typically are poorly sorted, and the grain sizes present range over three orders of magnitude—e.g., from 2 to 2,000 micrometres (8 × 10-5 to 8 × 10-2 inch). Commonly, the coarsest part of a wacke bed is its base, where pebbles may be abundant. Shale fragments, which represent lumps of mud eroded from bottom sediments by the depositing current, may be concentrated elsewhere in the bed.

      Many wackes contain much mud, typically 15–40 percent, and this increases as the mean grain size of the rock decreases. The particles forming the rock are typically angular. This, and the presence of the interstitial mud matrix, has led to these rocks being called “microbreccias.” The fabric and texture indicate that the sediments were carried only a short distance and were subject to very little reworking by currents after deposition.

      The most widespread internal structure of wackes is graded bedding, although some sequences display it poorly. Sets of cross strata more than three centimetres thick are rare, but thinner sets are very common. Parallel lamination is widespread, and convolute bedding is usually present. These internal structures are arranged within wacke beds in a regular sequence. They appear to result from the action of a single current flow and are related to changes in the hydraulics of the depositing current. In some beds, the upper part of the sequence of structures is missing, presumably because of erosion or nondeposition. In others, the lower part is missing. This has been attributed to change in the hydraulic properties of the depositing current as it moves away from its source and its velocity decreases to the point at which the first sediment deposited is laminated, rather than massive and graded as is the case closer to the source.

 The most typical external structures of wacke beds are sole markings, which occur on their undersurfaces. Flute and groove molds (see photographs—>) are the most characteristic, but many other structures have been recorded.

      The upper surfaces of wacke beds are less well characterized by sedimentary structures. The most typical are current lineation and various worm tracks, particularly of the highly sinuous form Nereites. Apart from these trace fossils, wackes are usually sparsely fossiliferous. Where fossils occur they are generally free-floating organisms (graptolites, foraminiferans) that have settled to the bottom, or bottom-living (benthic), shallow-water organisms displaced into deeper water as part of the sediment mass.

      Wackes are chemically homogeneous and are generally rich in aluminum oxide (Al2O3), ferrous oxide (FeO) + ferric oxide (Fe2O3), magnesium oxide (MgO), and soda (Na2O). The abundance of soda relative to potash (K2O) (reflecting a typically high sodium plagioclase feldspar content) and dominance of ferrous oxide over ferric oxide (reflecting large amounts of chlorite in the matrix) chemically distinguishes wackes from the three arenite families. The bulk composition of most wackes mimics that of their source owing to a lack of chemical differentiation by weathering and sorting. The matrix component, which is by definition any clasts 30 micrometres or finer, allows wackes to be differentiated from the other major sandstones. To be characterized as a wacke, its matrix component must equal or exceed 15 percent; in some cases more than 50 percent matrix has been reported. The origin of the matrix component, however, is controversial. Even though laboratory studies demonstrate that gravity-driven, bottom-hugging turbidity currents deposit sand-size grains together with mud-size clasts, modern deep-sea fan and abyssal plain sands (turbidites) have a matrix component that seldom exceeds 10 percent. A large portion of the matrix in ancient wackes must therefore be secondary, derived either from the disaggregation of feldspar and fine-grained lithic fragments like shale, phyllite, and volcanic rocks or from the postdepositional infiltration of clay- and silt-size clasts from overlying beds.

      Wackes are widespread in the geologic record and occur throughout geologic history. They typically are not found in association with sedimentary rocks that accumulate upon stable continental blocks and are instead confined either to intensely deformed mountain systems or to their modern analogues: ocean trenches, the continental slope and rise, and abyssal plain areas. Many, perhaps most, wackes are redeposited marine sands derived from source areas in which weathering, erosion, and deposition are too rapid to permit chemical differentiation and the breakdown of unstable components. Wackes of Archean age (those formed from 3.8 to 2.5 billion years ago) constitute the dominant sandstone type in the classic greenstone belts of the Precambrian shields (large areas of basement rocks in a craton that formed 3.8 billion to 570 million years ago around which younger sedimentary rocks have been deposited). They probably accumulated in rapidly subsiding trenches and ocean basins that surrounded primitive continental blocks. Proterozoic wackes (those formed from about 2.5 billion to 570 million years ago) are dominantly trench and ocean basin deposits, as are wackes of Phanerozoic age (those formed from 570 million years ago to the present day). They represent the accumulation of sand-size prisms of material that today are deposited both within ocean trenches (e.g., the modern trenches off Indonesia) and as submarine fan aprons (e.g., the Astoria Fan off the Pacific coast of Washington and Oregon in the United States) developed at the base of the continental slope at the mouths of submarine canyons. More distal carpets of wacke sand can extend for thousands of square kilometres across oceanic abyssal plains. Classic examples of the continental margin and ocean basin deposits include the late Precambrian Ocoee Supergroup and Ordovician Martinsburg Formation of the Appalachians, the Jurassic and Cretaceous Franciscan Formation of the Pacific Coast Ranges of California, much of the Alpine flysch (see below) of Switzerland and France, and many of the famous turbidite sands found in the Italian Apennines.

      The feature common to all modern depositional sites is that they adjoin landmasses in areas of high submarine relief. The landmass may be a continent bordered by either a passive, aseismic margin (for example, the eastern margin of North America) or a seismically active margin like that found along the western coast of both North and South America. The landmass can also be an active volcanic arc such as the Aleutian Islands chain or the Japan islands arc. The critical factor is the close proximity of topographically high and emergent clastic source areas and steeply sloped submarine depositional slopes, basins, or trenches.

Keith A.W. Crook Frederick L. Schwab

Mudrocks
      In terms of volume, mudrocks are by far the most important variety of sedimentary rock, probably constituting nearly 80 percent of the Earth's sedimentary rock column. Despite this abundance, the literature on mudrocks does not match in extent or detail that dealing with sandstones, carbonate rocks, and the various rarer sedimentary rock varieties like evaporite and phosphorite. This paradox reflects the difficulties inherent both in analyzing such rocks, owing to their poor exposure and fine grain size, and in interpreting any data obtained from their analysis because of the effects of diagenesis. Mudrocks include all siliciclastic sedimentary rocks composed of silt- and clay-size particles: siltstone (1/16 millimetre to 1/256 millimetre diameters), claystone (less than 1/256 millimetre), and mudstone (a mix of silt and clay). Shale refers specifically to mudrocks that regularly exhibit lamination or fissility or both. Mudrocks are also loosely referred to as both lutites and pelites and as argillaceous sedimentary rocks.

      Though mudrocks are composed mainly of detritus weathered from preexisting rocks, many contain large amounts of chemically precipitated cement (either calcium carbonate or silica), as well as abundant organic material. Mudrocks produced from the alteration of volcanic lava flows and ash beds to clay and zeolite minerals are called bentonites (bentonite).

General properties of shales (shale)
      The properties of shales are largely determined by the fine grain size of the constituent minerals. The accumulation of fine clastic detritus (silt) generally requires a sedimentary (sedimentation) environment of low mechanical energy (one in which wave and current actions are minimal), although some fine material may be trapped by plants or deposited as weakly coherent pellets in more agitated environments. The properties of the clay mineral constituents of lutites are particularly important, even when they do not make up the bulk of a rock.

      The mineralogy of shales is highly variable. In addition to clay minerals (60 percent), the average shale contains quartz and other forms of silica, notably amorphous silica and cristobalite (30 percent), feldspars (5 percent), and the carbonate minerals calcite and dolomite (5 percent). Iron oxides and organic matter (about 0.5 and 1 percent, respectively) are also important. Older estimates greatly underestimated clay minerals because of incorrect assignment of potassium to feldspar minerals. The most abundant clay mineral is illite; montmorillonite and mixed-layer illite-montmorillonite are next in abundance, followed by kaolinite, chlorite, chlorite-montmorillonite, and vermiculite. The quartz-to-feldspar ratio generally mirrors that of associated sands. In pelagic (deep-sea) sediments, however, feldspar may be derived from local volcanic sources, whereas quartz may be introduced from the continents by wind, upsetting simple patterns. A large number of accessory minerals occur in shales (stratification). Some of these are detrital, but diagenetic or in situ varieties (e.g., pyrite, siderite, and various phosphates) and volcanically derived varieties (e.g., zeolites, zircon, and biotite) have been noted.

Origin of shales
      The formation of fine-grained sediments generally requires weak transporting currents and a quiet depositional basin. Water is the common transporting medium, but ice-rafted glacial flour (silt produced by glacial grinding) is a major component in high-latitude oceanic muds, and windblown dust is prominent, particularly in the open ocean at low and intermediate latitudes. Shale environments thus include the deep ocean; the continental slope and rise; the deeper and more protected parts of shelves, shallow seas, and bays; coastal lagoons; interdistributory regions of deltas, swamps, and lakes (including arid basin playas); and river floodplains. The deep-sea muds are very fine, but an orderly sequence from coarse sediments in high-energy nearshore environments to fine sediments at greater depths is rarely found. Sediments at the outer edges of present-day continental shelves are commonly sands, relict deposits of shallower Pleistocene (from about 1,600,000 to 10,000 years ago) glacial conditions, whereas muds are currently being deposited in many parts of the inner shelf. The nearshore deposition of clay minerals is enhanced by the tendency of riverborne dispersed platelets to flocculate (flocculation) in saline waters (salinity greater than about four parts per thousand) and to be deposited just beyond the agitated estuarine environment as aggregates hydraulically equivalent to coarser particles. Differential flocculation leads to clay-mineral segregation, with illite and kaolinite near shore and montmorillonite farther out to sea. Advance of silty and sandy delta-slope deposits over clays also leads to complex grain-size patterns.

      Shales may be deposited in environments of periodic agitation. Sediments deposited on submarine slopes are frequently mechanically unstable and may be redistributed by slumping and turbidity currents to form thick accumulations (possible present-day eugeosynclinal equivalents) on the lower continental slope and rise. Part of the shale in many wacke-shale alternations may be of turbidite origin. Fine sediment can be deposited in marshes and on tidal flats. Trapping by marsh plants and binding of muds in fecal pellets are important. Because of electrochemical interactions among fine particles, muds plastered on a tidal flat by an advancing tide are difficult to reerode on the ebb. This may lead, as in the present-day Waddenzee, in The Netherlands, to a size increase from nearshore tidal flat muds to lag sands seaward. Fine floodplain sediments may dry out to coherent shale pellets, and these, on reerosion, can be redistributed as sands and gravels.

Shales of economic value
      Black shales are often of economic importance as sources of petroleum products and metals, and this importance will probably increase in the future. The lacustrine Eocene Green River Shales of Colorado, Wyoming, and Utah are potentially rich petroleum sources and are undergoing exploratory extraction. Bituminous layers of the Early Permian Irati Shales of Brazil are similarly important. These shales contain the remains of the marine reptile Mesosaurus, also found in South Africa, and played a prominent part in the development of the concepts of continental drift. The widespread thin Chattanooga Shale (Devonian-Mississippian) of the eastern United States has been exploited for its high (up to 250 parts per million) uranium content. The Kupferschiefer of the Permian (286 to 245 million years old) is a bituminous shale rich in metallic sulfides of primary sedimentary or early diagenetic origin; it covers a large area of central Europe as a band generally less than one metre thick, and in eastern Germany and in Poland there is sufficient enrichment in copper, lead, and zinc for its exploitation as an ore.

Kevin Charles Beck Frederick L. Schwab

Limestones (limestone) and dolomites
 Limestones and dolomites are collectively referred to as carbonates (carbonate rock) because they consist predominantly of the carbonate minerals calcite (CaCO3) and dolomite (CaMg[CO3]2). (See photographs—>.) Almost all dolomites are believed to be produced by recrystallization of preexisting limestones, although the exact details of this dolomitization process continue to be debated. Consequently, the following discussion initially deals with limestones and dolomites as a single rock type and subsequently considers the complex process by which some limestones become dolomite.

      Carbonates are by far the only volumetrically important nonsiliciclastic sedimentary rock type. Most are marine, and thick sequences of carbonate rocks occur in all the continental blocks, a surviving record of the transgressions and regressions of shallow marine (epeiric) seas that repeatedly blanketed the stable continental cratonic areas from time to time mainly during the late Precambrian, Paleozoic, and Mesozoic eras. Modern marine carbonate sediments, whose formation is favoured by warm, shallow water, are presently being deposited in a broad band straddling the Equator. The texture, sedimentary structures, composition, and organic content of carbonates provide numerous insights into the environment of deposition and regional paleogeography. Many important oil reservoirs of the world, especially those of the Middle East, occur in carbonate rocks.

      Though ancient limestones and dolomites are composed of calcite and dolomite, respectively, other calcite group minerals such as magnesite (MgCO3), rhodochrosite (MnCO3), and siderite (FeCO3) occur in limited amounts in restricted environments. Modern carbonate sediments are composed almost entirely of metastable aragonite (CaCO3) and magnesium-rich calcite, both of which readily recrystallize during diagenesis to form calcite. Carbonate rocks commonly grade naturally into siliciclastic sedimentary rocks as the proportion of terrigenous grains of varying size and mineralogy increases. Such mixtures are the consequence of the infringement of a dominantly siliciclastic depositional setting (e.g., a quartz arenitic beach area) into, for example, a lagoon or tidal flat in which carbonate mud accumulates.

Textural components
      Carbonate minerals present in ancient limestones and dolomites occur in one of three textural forms: (1) discrete silt to sand to coarser carbonate grains, or allochems, such as oöids or skeletal fragments, (2) mud-size interstitial calcium carbonate matrix called microcrystalline calcite or micrite, and (3) interlocking, 0.02- to 0.1-millimetre-diameter crystals of clear interstitial calcium carbonate cement or spar. In a rather simplistic sense, these three carbonate rock textural components are comparable, respectively, to the three possible constituents in a sandstone: (1) the coarser rock and mineral grains, (2) interstitial matrix, and (3) interstitial chemical cement.

      Several types of allochems exist: oöids, skeletal grains, carbonate clasts, and pellets. Oöids (also known as oölites or oöliths) are sand-size spheres of calcium carbonate mud concentrically laminated about some sort of nucleus grain, perhaps a fossil fragment or a silt-size detrital quartz grain. Oöids develop today on shallow shelf areas where strong bottom currents can wash the various kinds of material that form oöid nuclei back and forth in well-agitated, warm water that is supersaturated with calcium carbonate. The concentric layers of aragonite (in modern oöids) is produced by blue-green algae that affix themselves to the grain nucleus. Skeletal fragments, also known as bioclasts, can be whole fossils or broken fragments of organisms, depending on current and wave strength as well as depositional depth. The content and texture of the bioclast component in any carbonate will vary noticeably as a function of both age (due to evolution) and depositional setting (because of subsequent abrasion and transport as well as ecology). Carbonate clasts include fragments weathered from carbonate source rocks outside the depositional basin (lithoclasts) as well as fragments of carbonate sediment eroded from within the basin almost immediately after it was deposited (intraclasts). Silt- to sand-size particles of microcrystalline calcite or aragonite that lack the internal structure of oöids or bioclasts generally are called pellets or peloids. Most are fecal pellets generated by mud-ingesting organisms. Pellets can be cemented together into irregularly shaped composite grains dubbed lumpstones or grapestones.

      Microcrystalline carbonate mud ( micrite) and sparry carbonate cement (sparite) are collectively referred to as orthochemical carbonate because, in contrast to allochems, neither exhibits a history of transport and deposition as clastic material. Micrite can occur either as matrix that fills or partly fills the interstitial pores between allochems or as the main component of a carbonate rock. It originates mainly as the result of organic activity: algae generate tiny needles of aragonite within their tissues, and after their death such needles fall to the depositional surface as unconsolidated mud, which soon recrystallizes to calcite. Some micrite is produced by inorganic precipitation of aragonite; grain-to-grain collision and the resulting abrasion of allochems also can generate modest amounts of micrite. Most of the coarser and clearer crystals of sparry calcite that fill interstitial pores as cement represent either recrystallized micrite or essentially a direct inorganic precipitate.

      A number of carbonate classification schemes have been developed, but most modern ones subdivide and name carbonate rock types on the basis of the kinds of allochems present and the nature of the interstitial pore filling, whether it is micrite or spar. The most widely used scheme of this type is the descriptive classification devised by the American petrologist Robert L. Folk.

Origin of limestones
      Limestones originate mainly through the lithification of loose carbonate sediments. Modern carbonate sediments are generated in a variety of environments: continental, marine, and transitional, but most are marine. The present-day Bahama banks is the best known modern carbonate setting. It is a broad submarine shelf covered by shallow, warm seawater. The Bahama shelf, or carbonate platform, mimics the setting that repeatedly prevailed across the stable cratonic areas of the major continental blocks during late Precambrian, Paleozoic, and Mesozoic time and serves as a model for explaining the various limestone types that make up such ancient carbonate successions.

      The edge of the shelf is marked by a topographically sharp escarpment flanked by coarse, angular limestone breccia. Submarine channels etched into the escarpment serve as waterways down which shallow-water carbonate sediment can be transported by turbidity currents capable of redistributing them as apronlike deposits on the oceanic abyssal plain. In many areas, the fringe of the Bahama banks is marked by wave-resistant reef rocks (sometimes classified as boundstone). Abrasion of these reefs by wave activity generates abundant skeletal debris. Variations in depth and current strength control the relative amounts of micrite and sparite, the prevalence of specific organisms and their productivity, and the likelihood of generating oöids, pellets, and carbonate rock fragments. Micrite and micritic allochemical sediments accumulate in deep-water, low-energy, protected areas like lagoons and tidal flats and on the leeward side of major islands. In high-energy, shallow-water locales such as beaches, coastal dunes, and tidal channels, currents winnow out any micrite, and these become the sites of sparry allochemical sediment deposition. Pinpointing the exact depositional setting for an ancient carbonate deposit requires detailed analysis of its texture, composition, sedimentary structures, geometry, fossil content, and stratigraphic relationships with modern carbonate depositional sites.

      In addition to the ancient analogues of the modern carbonate deposits described above are freshwater limestones (marls (marl)) and limestone muds (or calcilutites) of deep-water abyssal plains (abyssal plain). Freshwater limestones of limited extent represent a spectrum of small-scale settings developed within and along the margins of lacustrine basins. Deep-water abyssal plain limestones are quite restricted in volume and age in the geologic record for a number of reasons. First of all, abyssal plain sequences are less likely to be incorporated into the orogenic belts that develop as continental margins are compressed during ocean basin closure. Second, pelagic calcareous oozes are the obvious modern analogues of ancient abyssal plain calcilutites. These oozes are produced by aragonite-secreting plankton that float near the surface (such as foraminiferans and coccoliths), which upon their death leave their shells, or tests, to settle slowly to the ocean bottom and accumulate. The development of such deep-sea deposits is therefore obviously dependent on the existence of calcium-secreting planktonic organisms, and these did not evolve until Mesozoic time. Finally, calcareous ooze accumulation is severely restricted both by latitude (being largely confined to a band extending 30° to 40° north and south of the Equator) and abyssal plain depth (approximately 2,000 metres). Below a depth of about 4,500 metres, which is the carbonate compensation depth (CCD), the pressure and temperature of seawater produces a rate of dissolution in excess of the rate of pelagic test accumulation.

Dolomites and dolomitization
      Dolomite is produced by dolomitization, a diagenetic process in which the calcium carbonate minerals aragonite and calcite are recrystallized and converted into the mineral dolomite. Dolomitization can obscure or even obliterate all or part of the original limestone textures and structures; in the case where such original features survive, carbonate nomenclature and interpretation can still be applied to the rock with emphasis on the effects of alteration.

      The exact processes by which limestones are dolomitized are not thoroughly understood, but dolomites occur widely in the geologic record. The relative proportion of dolomite to limestone progressively increases with age in carbonate rocks. This secular trend probably either reflects the earlier existence of geochemical settings that were more favourable to dolomitization or is the logical result of the fact that the likelihood for a limestone to undergo dolomitization increases proportionally with its age.

      Geochemists have been unable to precipitate normal dolomite under the conditions of temperature and pressure that exist in nature; temperatures within the 200° C range are required to support precipitation. A few modern, so-called primary marine dolomite localities have been studied, but close investigation of these areas suggests that even these penecontemporaneous dolomites are produced by altering calcite or aragonite almost immediately after their initial precipitation. Dolomites generated by later alteration of older limestones are known as diagenetic dolomites.

      The study of the few reported penecontemporaneous dolomite sites allows some conclusions to be formed regarding the dolomitization process. These modern dolomites develop mainly under conditions of high salinity (hypersalinity), which commonly exist in arid regions across supratidal mud flats as well as on the flat, saline plains and playa lake beds known as sabkhas. In highly saline environments, the ratio of dissolved magnesium ions to dissolved calcium ions progressively increases above the norm for seawater (5:1) as a result of the selective formation of calcium-rich evaporite minerals like gypsum and anhydrite. These magnesium-rich brines then tend to be flushed downward owing to their high density; the entire process is named evaporative reflux. Penecontemporaneous dolomites would result from the positioning of sabkhas and arid supratidal flats in a site that is in immediate contact with carbonate sediment; diagenetic dolomites would logically result when such dolomite-producing settings overlie older limestone deposits. The presence of fissures or highly permeable zones serving as channelways for downward percolation of dolomitizing fluids would also promote the alteration. Other studies have emphasized a possible role in dolomitization for dense brackish (salty) fluids formed when seawater and meteoric waters (those precipitated from the atmosphere as rain or snow) are produced along coastal zones.

Siliceous rocks (siliceous rock)
      Those siliceous rocks composed of an exceptionally high amount of crystalline siliceous material, mainly the mineral quartz (especially microcrystalline quartz and fibrous chalcedony) and amorphous opal, are most commonly known as chert (chert and flint). A wide variety of rock names are applied to cherty rocks reflecting their colour (flint is dark chert; jasper is usually red; prase is green) and geographic origin (novaculite of Arkansas, U.S.; silexite of France). The term chert is applied here to all fine-grained siliceous sediments and sedimentary rocks of chemical, biochemical, and organic origin.

Types of cherts
      Two major varieties of chert deposits exist—namely, bedded chert and nodular chert. Bedded cherts occur in individual bands or layers ranging in thickness from one to several centimetres or even tens of metres. They are intimately associated with volcanic rocks, commonly submarine volcanic flows as well as deep-water mudrocks. Classic examples include the Miocene Monterey Formation of the Coast Ranges of California, the Permian Rex Chert of Utah and Wyoming, the Arkansas Novaculite of the Ouachita Mountains, and the Mesozoic chert deposits of the Franciscan Formation of California. Nodular cherts occur as small to large (millimetres to centimetres) knotlike and fistlike clusters of quartz, chalcedony, and opal concentrated along or parallel with bedding planes in shallow-water marine carbonate rocks as well as pelagic limestones. Individual nodules may be ovoid or semispherical in shape; masses of chert typically form an anastamosing network.

Origin of cherts
      Many bedded cherts are composed almost entirely of the remains of silica-secreting organisms like diatoms and radiolarians. Such deposits are produced by compacting and recrystallizing the organically produced siliceous ooze deposits that accumulate on the present-day abyssal ocean floor. The modern oozes gather in latitudes where high organic productivity of floating planktonic radiolarians and diatoms takes place in the warm surface waters. As individual organisms die, their shells settle slowly to the abyssal floor and accumulate as unconsolidated siliceous ooze. Siliceous oozes are particularly prominent across areas of the ocean floor located far from continental blocks, where the rate of terrigenous sediment supply is low, and in deeper parts of the abyssal plain lying below the carbonate compensation depth, where the accumulation of calcareous oozes cannot occur. Some bedded cherts might not be of organic origin. They instead may be produced by precipitating silica gels derived from the same magma chambers from which the submarine basalts (pillow lava) that are intimately associated with bedded cherts are precipitated.

      The origin of nodular cherts has long been debated, but most are produced by the secondary replacement of the carbonate minerals and fossils within shallow marine shelf deposits. Evidence of secondary origin includes relict structures of allochems such as skeletal fragments and oöids preserved entirely within chert nodules. Silica can be mobilized from elsewhere within a rock and transported in solution under proper conditions of temperature and geochemistry. Likely sources of silica found scattered within shallow-water shelf carbonates include siliceous sponge spicules, radiolarians or diatom shells, and windblown sand grains. The details of the process and the possible role of microscopic organisms like bacteria in dissolving, mobilizing, and reconcentrating the silica remains uncertain.

      Finally, geysers and hot springs like those of the Yellowstone National Park area of northwestern Wyoming, U.S., are also sites of chert deposition. Encrustations of silica, known as sinter or geyserite, are volumetrically unimpressive but nevertheless are curiosities. The geyser and hot springs activity at Yellowstone is probably typical, with a subterranean body of magma as the source of silica-rich hydrothermal solutions rising periodically near or to the surface.

Phosphorites (phosphorite)
      Many sedimentary rocks contain phosphate in the form of scattered bones composed of the mineral apatite (calcium phosphate), but rocks composed predominantly of phosphate are rare. Nevertheless, three principal types exist: (1) regionally extensive, crystalline nodular, and bedded phosphorites, (2) localized concentrations of phosphate-rich clastic deposits (bone beds), and (3) guano deposits.

Evaporites (evaporite)
      Evaporites are layered crystalline sedimentary rocks that form from brines generated in areas where the amount of water lost by evaporation (vaporization) exceeds the total amount of water from rainfall and influx via rivers and streams. The mineralogy of evaporite rocks is complex, with almost 100 varieties possible, but less than a dozen species are volumetrically important. Minerals in evaporite rocks include carbonates (especially calcite, dolomite, magnesite, and aragonite), sulfates (anhydrite and gypsum), and chlorides (particularly halite, sylvite, and carnallite), as well as various borates, silicates, nitrates, and sulfocarbonates. Evaporite deposits occur in both marine and nonmarine sedimentary successions.

      Though restricted in area, modern evaporites contribute to genetic models for explaining ancient evaporite deposits. Modern evaporites are limited to arid regions (those of high temperature and low rates of precipitation), for example, on the floors of semidry ephemeral playa lakes in the Great Basin of Nevada and California, across the coastal salt flats (sabkhas) of the Middle East, and in salt pans, estuaries, and lagoons around the Gulf of Suez. Ancient evaporates occur widely in the Phanerozoic geologic record, particularly in those of Cambrian (from 570 to 505 million years ago), Permian (from 286 to 245 million years ago), and Triassic (from 245 to 208 million years ago) age, but are rare in sedimentary sequences of Precambrian age. They tend to be closely associated with shallow marine shelf carbonates and fine (typically rich in iron oxide) mudrocks. Because evaporite sedimentation requires a specific climate and basin setting, their presence in time and space clearly constrains inferences of paleoclimatology and paleogeography. Evaporite beds tend to concentrate and facilitate major thrust fault horizons, so their presence is of particular interest to structural geologists. Evaporites also have economic significance as a source of salts and fertilizer.

      All evaporite deposits result from the precipitation of brines (brine) generated by evaporation. Laboratory experiments can accurately trace the evolution of brines as various evaporite minerals crystallize. Normal seawater has a salinity of 3.5 percent (or 35,000 parts per million), with the most important dissolved constituents being sodium and chlorine. When seawater volume is reduced to one-fifth of the original, evaporite precipitation commences in an orderly fashion, with the more insoluble components (gypsum and anhydrite) forming first. When the solution reaches one-tenth the volume of the original, more soluble minerals like sylvite and halite form. Natural evaporite sequences show vertical changes in mineralogy that crudely correspond to the orderly appearance of mineralogy as a function of solubility but are less systematic.

Nonmarine environment
      Evaporite deposition in the nonmarine environment occurs in closed lakes—i.e., those without outlet—in arid and semiarid regions. Such lakes form in closed interior basins or shallow depressions on land where drainage is internal and runoff does not reach the sea. If water depths are shallow or, more typically, somewhat ephemeral, the term playa or playa lake is commonly used.

      Water inflow into closed lakes consists principally of precipitation and surface runoff, both of which are small in amount and variable in occurrence in arid regions. Groundwater flow and discharge from springs may provide additional water input, but evaporation rates are always in excess of precipitation and surface runoff. Sporadic or seasonal storms may give rise to a sudden surge of water inflow. Because closed lakes lack outlets, they can respond to such circumstances only by deepening and expanding. Subsequent evaporation will reduce the volume of water present to prestorm or normal amount; fluctuation of closed lake levels therefore characterizes the environment.

      Such changing lake levels and water volumes lead to fluctuating salinity values. Variations in salinity effect equilibrium relations between the resulting brines and lead to much solution and subsequent reprecipitation of evaporites in the nonmarine environment. As a result of these complexities as well as the distinctive nature of dissolved constituents in closed lake settings, nonmarine evaporite deposits contain many minerals that are uncommon in marine evaporites—e.g., borax, epsomite, trona, and mirabilite.

Shallow marine environment
      Evaporite deposition in the shallow marine environment (sometimes termed the salina (alkali flat)) occurs in desert coastal areas, particularly along the margins of such semi-restricted water bodies as the Red Sea, Persian Gulf, and Gulf of California. Restriction is, in general, one of the critical requirements for evaporite deposition, because free and unlimited mixing with the open sea would allow the bodies of water to easily overcome the high evaporation rates of arid areas and dilute these waters to near-normal salinity. This semi-restriction cannot, in fact, prevent a large amount of dilution by mixing; coastal physiography is the principal factor involved in brine production. Shallow-water evaporites, almost exclusively gypsum, anhydrite, and halite, typically interfinger with tidal flat limestone and dolomite and fine-grained mudrock.

Deep-basin environment
      Most of the thick, laterally extensive evaporite deposits appear to have been produced in deep, isolated basins that developed during episodes of global aridity. The most crucial requirement for evaporite production is aridity; water must be evaporated more rapidly than it can be replenished by precipitation and inflow. In addition, the evaporite basin must somehow be isolated or at least partially isolated from the open ocean so that brines produced through evaporation are prevented from returning there. Restricting brines to such an isolated basin over a period of time enables them to be concentrated to the point where evaporite mineral precipitation occurs. Periodic breaching of the barrier, due either to crustal downwarping or to global sea-level changes, refills the basin from time to time, thereby replenishing the volume of seawater to be evaporated and making possible the inordinately thick, regionally extensive evaporite sequences visible in the geologic record.

 Debate continues over the exact mechanisms for generating thick evaporite deposits. Three possible models for restricting “barred” evaporite basins are shown in Figure 5—>. They differ in detail, and none has garnered a consensus of support. The deep-water, deep-basin model accounts for replenishment of the basin across the barrier or sill, with slow, continual buildup of thick evaporites made possible by the seaward escape of brine that allows a constant brine concentration to be maintained. The shallow-water, shallow-basin model produces thick evaporites by continual subsidence of the basin floor. The shallow-water, deep-basin model shows the brine level in the basin beneath the level of the sea as a result of evaporation; brines are replenished by groundwater recharge from the open ocean.

Iron-rich (iron) sedimentary rocks
      Almost all sedimentary rocks are iron-bearing in the sense that mudrocks, sandstones, and carbonates typically have an iron content of several percent. Nevertheless, sedimentary rocks in which the proportion of iron exceeds 15 percent are separately categorized as iron-rich. Two major types of iron-rich sedimentary rocks are recognized: (1) iron formation, or banded iron formation (BIF (banded-iron formation))—regionally extensive, locally thick sequences composed of alternating thin (millimetre to centimetre thick) layers of mainly crystalline-textured iron-rich minerals and chert—and (2) ironstone—noncherty, essentially clastic-textured, iron-rich minerals of local extent.

      Banded iron formations are predominantly Precambrian in age; most are 1.8 to 2.2 billion years old; none are younger than Cambrian age. The most important iron-bearing minerals in iron formations are hematite, magnetite, and greenalite. These deposits constitute the world's major source of iron ore. Classic examples are found in the Mesabi Range of Minnesota, U.S., and the Kiruna ores of Sweden.

      Ironstones are principally of Phanerozoic age, mainly Early Paleozoic (roughly 440 to 570 million years old) and Jurassic (about 144 to 208 million years old), but can be as old as Middle Precambrian age (about 1.6 to 3 billion years old). They appear to be restricted to basins no larger than 150 kilometres in any direction. Major iron minerals are goethite, hematite, and chamosite.

Origin of banded iron formation
      The origin of banded iron formation is not clearly understood. Banded iron formation units are typically 50 to 600 metres thick. Their complex mineralogy includes various iron oxides, iron carbonates, iron silicates, and iron sulfides. The essentially crystalline texture of these minerals together with the definitive crystalline texture of the laminated chert bands with which the iron mineral layers alternate is perplexing. At present, iron is not easily dissolved, nor can it be readily transported in solution and subsequently precipitated as crystalline-textured, iron-rich minerals, because of the presence of free atmospheric oxygen. Many sedimentary petrologists consequently conclude that banded iron formation deposition is a uniquely Precambrian occurrence made possible by, and supporting the existence of, an earlier anaerobic Earth atmosphere (one lacking free oxygen) quite unlike that in existence today. Controversy also continues over the ultimate source of iron (weathering as opposed to magmatic iron escaping from the Earth's interior) and over the possible role of microorganisms such as bacteria and algae in the precipitation of the iron.

Origin of ironstones
      The origin of ironstones also is not well understood, but most appear to be derived from the erosion and redeposition of lateritic (iron-rich, red) soils. Ironstones occur as thin (a few tens of metres at most) units interbedded with shallow marine and transitional carbonates, mudrocks, and sandstones. They generally have an oölitic texture. Cross-bedding, ripple marks, and small scour and fill channels are abundant. Slight uplift and erosion of reddish soils developed in coastal regions drained by rivers that transport and deposit such material in deltas and embayments along the coast is compatible with the features and fossils found in deposits of this sort. Classical ironstone deposits include the Ordovician Wabana Formation of Newfoundland and the Silurian Clinton Group of the central and southern Appalachians.

Organic-rich sedimentary deposits (fossil fuel)
      Coal, oil shale, and petroleum are not sedimentary rocks per se; they represent accumulations of undecayed organic tissue that can either make up the bulk of the material (e.g., coal), or be disseminated in the pores within mudrocks, sandstones, and carbonates (e.g., oil shale and petroleum). Much of the undecomposed organic matter in sediment and sedimentary rocks is humus, plant matter that accumulates in soil. Other important organic constituents include peat, humic organic matter that collects in bogs and swamps where oxidation and bacterial decay is incomplete, and sapropel, fine-grained organic material—mainly the soft organic tissue of phytoplankton and zooplankton, along with bits and pieces of higher plants—that amass subaqueously in lakes and oceans.

      Organic-rich sedimentary rock deposits are collectively referred to as fossil fuels because they consist of the undecayed organic tissue of plants and animals preserved in depositional settings characterized by a lack of free oxygen. Fossil fuels constitute the major sources of energy in the industrial world, and their unequal distribution in time (exclusively Phanerozoic) and space (more than half of the proved petroleum reserves are in the Persian Gulf region of the Middle East) has a significant effect on the world's political and economic stability.

      Coals are the most abundant organic-rich sedimentary rock. They consist of undecayed organic matter that either accumulated in place or was transported from elsewhere to the depositional site. The most important organic component in coal is humus. The grade or rank of coal is determined by the percentage of carbon present. The term peat is used for the uncompacted plant matter that accumulates in bogs and brackish swamps. With increasing compaction and carbon content, peat can be transformed into the various kinds of coal: initially brown coal or lignite, then soft or bituminous coal, and finally, with metamorphism, hard or anthracite coal. In the geologic record, coal occurs in beds, called seams, which are blanketlike coal deposits a few centimetres to metres or hundreds of metres thick.

      Many coal seams occur within cyclothems (cyclothem), rhythmic successions of sandstone, mudrock, and limestone in which nonmarine units are regularly and systematically overlain by an underclay, the coal seam itself, and then various marine lithologies. The nonmarine units are thought to constitute the floor of ancient forests and swamps developed in low-lying coastal regions; the underclay is a preserved relict of the soil in which the coal-producing vegetation was rooted; and the marine units overlying the coal record the rapid transgression of the sea inland that killed the vegetation by drowning it and preventing its decomposition by rapid burial. The exact mechanism responsible for generating the rapid episodes of marine transgression and regression necessary to generate coal-bearing cyclothems is not definitively known. A combination of episodic upwarping (epeirogeny) and downwarping of the continental blocks or global (eustatic) changes in sea level or both, coupled with normal changes in the rate of sediment supply that occurs along coasts traversed by major laterally meandering river systems, may have been the cause.

      In any case, coal is a rare, though widely distributed, lithology. Extensive coal deposits overall occur mainly in rocks of Devonian age (those from 408 to 360 million years old) and younger because their existence is clearly contingent on the evolution of land plants. Nevertheless, small, scattered coal deposits as old as early Proterozoic have been described. Coal-bearing cyclothem deposits are especially abundant in the middle and late Paleozoic sequences of the Appalachians and central United States and in the Carboniferous of Great Britain, probably because during this time interval global climates were warm and humid and large portions of the continental blocks were low-lying platforms located only slightly above sea level.

      Mudrock containing high amounts of organic matter in the form of kerogen is known as oil shale or kerogen shale. Kerogen is altered, undecayed, fatty organic matter, mainly sapropel, and is very fine-grained. It can be generated in place or transported. The most famous oil shale deposit in the world, located in the United States, is the Green River Formation of Utah, Wyoming, and Colorado of Eocene age (i.e., formed 57.8 to 36.6 million years ago). This vast deposit contains fossils and sedimentary structures, suggesting rapid deposition and burial of unoxidized organic matter in shallow lakes or marine embayments. The quantity of oil entrapped in the Green River Formation is significant. The cost of extracting the oil trapped in the mudrock by heating the oil shale, however, far exceeds the cost of extracting equivalent quantities of oil, natural gas, or coal. Also, the refractory process not only requires extensive strip mining and immense volumes of water but also results in a volume expansion of the original oil shale. Despite these considerable economic and technical problems, oil (petroleum) shales potentially represent a significant future energy resource.

Oil and natural gas
      Natural gas refers collectively to the various gaseous hydrocarbons generated below the Earth's surface and trapped in the pores of sedimentary rocks. Major natural gas varieties include methane, ethane, propane, and butane. These natural gases are commonly, though not invariably, intimately associated with the various liquid hydrocarbons—mainly liquid paraffins, napthenes, and aromatics—that collectively constitute oil.

      Hydrocarbons (hydrocarbon) can also exist in a semisolid or solid state such as asphalt, asphaltites, mineral waxes, and pyrobitumens. Bitumens (bitumen) can occur as seepages, impregnations filling the pore space of sediments (e.g., tar sands of the Canadian Rocky Mountains), and in veins or dikes. Asphaltites (asphaltite) occur primarily in dikes and veins that cross sedimentary rocks such as gilsonite deposits in the Green River Formation of Utah. These natural bitumens probably form from the loss of volatiles, oxidation, and biological degradation resulting from oil seepage to the surface. Solid hydrocarbons are of interest to geologists as their presence is a good indicator of petroleum below the surface in that region. Also, solid hydrocarbons have commercial value.

      The exact process by which oil and natural gas are produced is not precisely known, despite the extensive efforts made to determine the mode of petroleum genesis. Crude oil is thought to form from undecomposed organic matter, principally single-celled floating phytoplankton and zooplankton that settle to the bottom of marine basins and are rapidly buried within sequences of mudrock and limestone. Natural gas and oil are generated from such source rocks only after heating and compaction. Typical petroleum formation (maturation) temperatures do not exceed 100° C, meaning that the depth of burial of source rocks cannot be greater than a few kilometres. After their formation, oil and natural gas migrate from source rocks to reservoir rocks composed of sedimentary rocks largely as a consequence of the lower density of the hydrocarbon fluids and gases. Good reservoir rocks, by implication, must possess high porosity and permeability. A high proportion of open pore spaces enhances the capacity of a reservoir to store the migrating petroleum; the interconnectedness of the pores facilitates the withdrawal of the petroleum once the reservoir rock is penetrated by drill holes.

Secular trends in the sedimentary rock record
      Reexamination of the sedimentary rock record preserved within the continental blocks suggests systematic changes through time in the relative proportions of the major sedimentary rock types deposited, as summarized in Figure 6. These changes can be linked to the evolution of the atmosphere and hydrosphere and to the changing global tectonic setting. Carbonates and quartz sands, for example, require long-term source area stability as well as the existence of broad, shallow-water epeiric seas that mantle continental blocks. Marine transgressions and regressions across broad stable continental cratons occurred only in Proterozoic and Phanerozoic time; Archean continental blocks were smaller and tectonically unstable, and most likely less granitic than those of today. Consequently, the early Precambrian sedimentary rock record consists largely of volcanogenic sediments, wackes, and arkoses physically disintegrated from small, high-relief island arcs (the Archean greenstone belts of the various Precambrian shields) and microcontinental fragments. The fact that iron formations are restricted to rocks of Archean and Proterozoic time supports the conclusion that atmospheric oxygen levels in earlier stages of Earth history were lower, promoting the dissolution, transport, and precipitation of iron by chemical or biochemical means. The total lack of evaporites from the Archean record and their subsequent steady buildup probably reflects a number of factors. Such deposits can be easily destroyed by metamorphism, and presumably, given enough time, they will have been completely erased. Also, evaporite formation requires seawater with elevated salinity, a condition that is established with time. Finally, significant volumes of bedded evaporites can occur only once broad, stable continents have evolved, because the requisite restricted evaporite basins develop exclusively adjacent to cratons.

Frederick L. Schwab

Additional Reading
Sam Boggs, Jr., Principles of Sedimentology and Stratigraphy, 3rd ed. (2001); Rhodes W. Fairbridge and Joanne Bourgeois, The Encyclopedia of Sedimentology (1978); Andrew D. Miall, Principles of Sedimentary Basin Analysis, 2nd ed. (1990); F.J. Pettijohn, Sedimentary Rocks, 3rd ed. (1975); F.J. Pettijohn, P.E. Potter, and R. Siever, Sand and Sandstone, 2nd ed. (1987); P.E. Potter, J. Barry Maynard, and Wayne A. Pryor, Sedimentology of Shale (1980); Donald R. Prothero, Interpreting the Stratigraphic Record (1989); J.G. Reading (ed.), Sedimentary Environments and Facies, 2nd ed. (1986); Stuart R. Taylor and Scott M. McLennan, The Continental Crust: Its Composition and Evolution: An Examination of the Geochemical Record Preserved in Sedimentary Rocks (1985); Maurice E. Tucker and V.P. Wright, Carbonate Sedimentology (1990); Maurice E. Tucker, Sedimentary Petrology: An Introduction to the Origin of Sedimentary Rocks, 3rd ed. (2001); James L. Wilson, Carbonate Facies in Geologic History (1975).Frederick L. Schwab

* * *


Universalium. 2010.

Игры ⚽ Нужен реферат?

Look at other dictionaries:

  • Sedimentary rock — Middle Triassic marginal marine sequence of siltstones (below) and limestones (above), Virgin Formation, southwestern Utah, USA Sedimentary rock are types of rock that are formed by the deposition of material at the Earth s surface and within… …   Wikipedia

  • sedimentary rock — nuosėdinė uoliena statusas T sritis chemija apibrėžtis Uoliena, susidariusi iš sausumoje arba vandens baseine nusėdusių medžiagų. atitikmenys: angl. sedimentary rock rus. осадочная порода …   Chemijos terminų aiškinamasis žodynas

  • sedimentary rock —    Rock formed by the deposit of eroded igneous rock, often in strata or beds. Examples include limestone and sandstone …   Glossary of Art Terms

  • sedimentary rock — noun rock formed from consolidated clay sediments • Hypernyms: ↑rock, ↑stone • Hyponyms: ↑argillite, ↑arenaceous rock, ↑argillaceous rock, ↑rudaceous rock, ↑slate, ↑ …   Useful english dictionary

  • sedimentary rock —    A consolidated deposit of clastic particles, chemical precipitates, or organic remains accumulated at or near the surface of the earth under normal low temperature and pressure conditions.    Sedimentary rocks include consolidated equivalents… …   Glossary of landform and geologic terms

  • sedimentary rock — noun one of the major groups of rock that makes up the crust of the Earth; formed by the deposition of either the weathered remains of other rocks, the results of biological activity, or precipitation from solution …   Wiktionary

  • sedimentary rock —   any rock formed from sediments under pressure from other overlying sediments …   Geography glossary

  • sedimentary rock — type of stone having a layered structure and created by a buildup of underwater sediment …   English contemporary dictionary

  • Rock microstructure — includes the texture of a rock and the small scale rock structures. The words texture and microstructure are interchangeable, with the latter preferred in modern geological literature. However, texture is still acceptable because it is a useful… …   Wikipedia

  • sedimentary — (adj.) 1764, from SEDIMENT (Cf. sediment) + ARY (Cf. ary). Sedimentary rock attested by 1830 (in Lyell) …   Etymology dictionary

Share the article and excerpts

Direct link
Do a right-click on the link above
and select “Copy Link”